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 843 毫秒
1.
A statistical method to predict protein pKa has been developed by using the 3D structure of a protein and a database of 434 experimental protein pKa values. Each pKa in the database is associated with a fingerprint that describes the chemical environment around an ionizable residue. A computational tool, MoKaBio, has been developed to identify automatically ionizable residues in a protein, generate fingerprints that describe the chemical environment around such residues, and predict pKa from the experimental pKa values in the database by using a similarity metric. The method, which retrieved the pKa of 429 of the 434 ionizable sites in the database correctly, was crossvalidated by leave‐one‐out and yielded root mean square error (RMSE) = 0.95, a result that is superior to that obtained by using the Null Model (RMSE 1.07) and other well‐established protein pKa prediction tools. This novel approach is suitable to rationalize protein pKa by comparing the region around the ionizable site with similar regions whose ionizable site pKa is known. The pKa of residues that have a unique environment not represented in the training set cannot be predicted accurately, however, the method offers the advantage of being trainable to increase its predictive power. Proteins 2009. © 2009 Wiley‐Liss, Inc.  相似文献   

2.
Fluorescent proteins homologous to green fluorescent protein (avGFP) display pronounced spectral variability due to different chromophore structures and variable chromophore interactions with the surrounding amino acids. To gain insight into the structural basis for yellow emission, the 3D structure of phiYFP (λem = 537 nm), a protein from the sea medusa Phialidium sp., was built by a combined homology modeling – mass spectrometry approach. Mass spectrometry of the isolated chromophore-bearing peptide reveals that the chromophore of phiYFP is chemically identical to that of avGFP (λem = 508 nm). The experimentally acquired chromophore structure was combined with the homology-based model of phiYFP, and the proposed 3D structure was used as a starting point for identification of the structural features responsible for yellow fluorescence. Mutagenesis of residues in the local chromophore environment of phiYFP suggests that multiple factors cooperate to establish the longest-wavelength emission maximum among fluorescent proteins with an unmodified GFP-like chromophore.  相似文献   

3.
Sidechain rotamer libraries are obtained through exhaustive statistical analysis of existing crystallographic structures of proteins and have been applied in multiple aspects of structural biology, for example, crystallography of relatively low‐resolution structures, in homology model building and in biomolecular NMR. Little is known, however, about the driving forces that lead to the preference or suitability of one rotamer over another. Construction of 3D hydropathic interaction maps for nearly 30,000 tyrosines reveals the environment around each, in terms of hydrophobic (π–π stacking, etc.) and polar (hydrogen bonding, etc.) interactions. After partitioning the tyrosines into backbone‐dependent (?, ψ) bins, a map similarity metric based on the correlation coefficient was applied to each map‐map pair to build matrices suitable for clustering with k‐means. The first bin (?200° ≤ ? < –155°; ?205° ≤ ψ < –160°), representing 631 tyrosines, reduced to 14 unique hydropathic environments, with most diversity arising from favorable hydrophobic interactions with many different residue partner types. Polar interactions for tyrosine include surprisingly ubiquitous hydrogen bonding with the phenolic OH and a handful of unique environments surrounding the tyrosine backbone. The memberships of all but one of the 14 environments are dominated (>50%) by a single χ1/χ2 rotamer. The last environment has weak or no interactions with the tyrosine ring and its χ1/χ2 rotamer is indeterminate, which is consistent with it being composed of mostly surface residues. Each tyrosine residue attempts to fulfill its hydropathic valence and thus, structural water molecules are seen in a variety of roles throughout protein structure. Proteins 2015; 83:1118–1136. © 2015 Wiley Periodicals, Inc.  相似文献   

4.
Bio3D is a family of R packages for the analysis of biomolecular sequence, structure, and dynamics. Major functionality includes biomolecular database searching and retrieval, sequence and structure conservation analysis, ensemble normal mode analysis, protein structure and correlation network analysis, principal component, and related multivariate analysis methods. Here, we review recent package developments, including a new underlying segregation into separate packages for distinct analysis, and introduce a new method for structure analysis named ensemble difference distance matrix analysis (eDDM). The eDDM approach calculates and compares atomic distance matrices across large sets of homologous atomic structures to help identify the residue wise determinants underlying specific functional processes. An eDDM workflow is detailed along with an example application to a large protein family. As a new member of the Bio3D family, the Bio3D‐eddm package supports both experimental and theoretical simulation‐generated structures, is integrated with other methods for dissecting sequence‐structure–function relationships, and can be used in a highly automated and reproducible manner. Bio3D is distributed as an integrated set of platform independent open source R packages available from: http://thegrantlab.org/bio3d/ .  相似文献   

5.
The atomic-level structural properties of proteins, such as bond lengths, bond angles, and torsion angles, have been well studied and understood based on either chemistry knowledge or statistical analysis. Similar properties on the residue-level, such as the distances between two residues and the angles formed by short sequences of residues, can be equally important for structural analysis and modeling, but these have not been examined and documented on a similar scale. While these properties are difficult to measure experimentally, they can be statistically estimated in meaningful ways based on their distributions in known proteins structures. Residue-level structural properties including various types of residue distances and angles are estimated statistically. A software package is built to provide direct access to the statistical data for the properties including some important correlations not previously investigated. The distributions of residue distances and angles may vary with varying sequences, but in most cases, are concentrated in some high probability ranges, corresponding to their frequent occurrences in either α-helices or β-sheets. Strong correlations among neighboring residue angles, similar to those between neighboring torsion angles at the atomic-level, are revealed based on their statistical measures. Residue-level statistical potentials can be defined using the statistical distributions and correlations of the residue distances and angles. Ramachandran-like plots for strongly correlated residue angles are plotted and analyzed. Their applications to structural evaluation and refinement are demonstrated. With the increase in both number and quality of known protein structures, many structural properties can be derived from sets of protein structures by statistical analysis and data mining, and these can even be used as a supplement to the experimental data for structure determinations. Indeed, the statistical measures on various types of residue distances and angles provide more systematic and quantitative assessments on these properties, which can otherwise be estimated only individually and qualitatively. Their distributions and correlations in known protein structures show their importance for providing insights into how proteins may fold naturally to various residue-level structures.  相似文献   

6.
Predicted protein residue–residue contacts can be used to build three‐dimensional models and consequently to predict protein folds from scratch. A considerable amount of effort is currently being spent to improve contact prediction accuracy, whereas few methods are available to construct protein tertiary structures from predicted contacts. Here, we present an ab initio protein folding method to build three‐dimensional models using predicted contacts and secondary structures. Our method first translates contacts and secondary structures into distance, dihedral angle, and hydrogen bond restraints according to a set of new conversion rules, and then provides these restraints as input for a distance geometry algorithm to build tertiary structure models. The initially reconstructed models are used to regenerate a set of physically realistic contact restraints and detect secondary structure patterns, which are then used to reconstruct final structural models. This unique two‐stage modeling approach of integrating contacts and secondary structures improves the quality and accuracy of structural models and in particular generates better β‐sheets than other algorithms. We validate our method on two standard benchmark datasets using true contacts and secondary structures. Our method improves TM‐score of reconstructed protein models by 45% and 42% over the existing method on the two datasets, respectively. On the dataset for benchmarking reconstructions methods with predicted contacts and secondary structures, the average TM‐score of best models reconstructed by our method is 0.59, 5.5% higher than the existing method. The CONFOLD web server is available at http://protein.rnet.missouri.edu/confold/ . Proteins 2015; 83:1436–1449. © 2015 Wiley Periodicals, Inc.  相似文献   

7.
Recent progress in magic-angle spinning (MAS) solid-state NMR (SSNMR) has enabled multidimensional studies of large, macroscopically unoriented membrane proteins with associated lipids, without the requirement of solubility that limits other structural techniques. Here we present initial sample preparation and SSNMR studies of a 144 kDa integral membrane protein, E. coli cytochrome bo3 oxidase. The optimized protocol for expression and purification yields ∼5 mg of the enzymatically active, uniformly 13C,15N-enriched membrane protein complex from each liter of growth medium. The preparation retains endogenous lipids and yields spectra of high sensitivity and resolution, consistent with a folded, homogenous protein. Line widths of isolated signals are less than 0.5 ppm, with a large number of individual resonances resolved in the 2D and 3D spectra. The 13C chemical shifts, assigned by amino acid type, are consistent with the secondary structure previously observed by diffraction methods. Although the structure is predominantly helical, the percentage of non-helical signals varies among residue types; these percentages agree well between the NMR and diffraction data. Samples show minimal evidence of degradation after several weeks of NMR data acquisition. Use of a triple resonance scroll resonator probe further improves sample stability and enables higher power decoupling, higher duty cycles and more advanced 3D experiments to be performed. These initial results in cytochrome bo3 oxidase demonstrate that multidimensional MAS SSNMR techniques have sufficient sensitivity and resolution to interrogate selected parts of a very large uniformly 13C,15N-labeled membrane protein. Heather L. Frericks and Lai Lai Yap contributed equally to this work.  相似文献   

8.
The structural annotation of proteins with no detectable homologs of known 3D structure identified using sequence‐search methods is a major challenge today. We propose an original method that computes the conditional probabilities for the amino‐acid sequence of a protein to fit to known protein 3D structures using a structural alphabet, known as “Protein Blocks” (PBs). PBs constitute a library of 16 local structural prototypes that approximate every part of protein backbone structures. It is used to encode 3D protein structures into 1D PB sequences and to capture sequence to structure relationships. Our method relies on amino acid occurrence matrices, one for each PB, to score global and local threading of query amino acid sequences to protein folds encoded into PB sequences. It does not use any information from residue contacts or sequence‐search methods or explicit incorporation of hydrophobic effect. The performance of the method was assessed with independent test datasets derived from SCOP 1.75A. With a Z‐score cutoff that achieved 95% specificity (i.e., less than 5% false positives), global and local threading showed sensitivity of 64.1% and 34.2%, respectively. We further tested its performance on 57 difficult CASP10 targets that had no known homologs in PDB: 38 compatible templates were identified by our approach and 66% of these hits yielded correctly predicted structures. This method scales‐up well and offers promising perspectives for structural annotations at genomic level. It has been implemented in the form of a web‐server that is freely available at http://www.bo‐protscience.fr/forsa .  相似文献   

9.
The 50‐residue snake venom protein L ‐omwaprin and its enantiomer D ‐omwaprin were prepared by total chemical synthesis. Radial diffusion assays were performed against Bacillus megaterium and Bacillus anthracis; both L ‐ and D ‐omwaprin showed antibacterial activity against B. megaterium. The native protein enantiomer, made of L ‐amino acids, failed to crystallize readily. However, when a racemic mixture containing equal amounts of L ‐ and D ‐omwaprin was used, diffraction quality crystals were obtained. The racemic protein sample crystallized in the centrosymmetric space group P21/c and its structure was determined at atomic resolution (1.33 Å) by a combination of Patterson and direct methods based on the strong scattering from the sulfur atoms in the eight cysteine residues per protein. Racemic crystallography once again proved to be a valuable method for obtaining crystals of recalcitrant proteins and for determining high‐resolution X‐ray structures by direct methods.  相似文献   

10.
BmK M4 is a neutral neurotoxin in the BmK toxin series. It is medially toxic and belongs to group III cc-toxins. The purified sample was crystallized in rhombic space group P6 Using an X-ray diffraction technique, the crystal structure of BmK M4 was revealed by molecular replacement at 0.20 nm resolution. The model was refined. The final crystallographic R factor was 0.142 and the free R factor was 0.173. The root mean square deviation is 0.001 5 nm for the bond length and 1.753° for the bond angles. 64 water molecules were added to the asymmetric unit. The refined structure showed an unusual non-prolyl cis peptide bond at residue 10. The structure was compared with group II a-toxin BmK M8 (an acidic, weak toxin). The potential structural implications of the cis peptide bond were discussed.  相似文献   

11.
We introduce a new method for assessing the extent of residue exposure in proteins. For each atom of every residue a Gaussian-weighted atomic surroundings value (the G-neighborhood) is calculated. A normalized sum of G-neighborhood values over all the atoms of a residue is complementary to conventional surface accessibility characteristics. The G-0neighborhood value of a residue is a sensitive indicator of its location, strongly dependent on the 3D structure of a the protein. Correlations between secondary structures and patterns of G-neighborhood values for six different protein molecules are discussed. Comparison of the distribution of hydrophobic and charged residues in the 3D structure for the alcohol-soluble protein crambin and that of five water-soluble proteins (cytochrome c, flavodoxin, myoglobin, rhodanese, and Bence–Jones protein) shows striking differences in their G-neighborhood patterns. Contacts between the prosthetic group and the peptide portion of a protein as well as protein interdomain contacts and monomer–monomer contacts are characterized. © 1993 Wiley-Liss, Inc.  相似文献   

12.
The elucidation of the mutual influence between peptide bond geometry and local conformation has important implications for protein structure refinement, validation, and prediction. To gain insights into the structural determinants and the energetic contributions associated with protein/peptide backbone plasticity, we here report an extensive analysis of the variability of the peptide bond angles by combining statistical analyses of protein structures and quantum mechanics calculations on small model peptide systems. Our analyses demonstrate that all the backbone bond angles strongly depend on the peptide conformation and unveil the existence of regular trends as function of ψ and/or φ. The excellent agreement of the quantum mechanics calculations with the statistical surveys of protein structures validates the computational scheme here employed and demonstrates that the valence geometry of protein/peptide backbone is primarily dictated by local interactions. Notably, for the first time we show that the position of the Hα hydrogen atom, which is an important parameter in NMR structural studies, is also dependent on the local conformation. Most of the trends observed may be satisfactorily explained by invoking steric repulsive interactions; in some specific cases the valence bond variability is also influenced by hydrogen‐bond like interactions. Moreover, we can provide a reliable estimate of the energies involved in the interplay between geometry and conformations. Proteins 2015; 83:1973–1986. © 2015 Wiley Periodicals, Inc.  相似文献   

13.
Canonical loops of protein inhibitors of serine proteinases occur in proteins having completely different folds. In this article, conformations of the loops have been analyzed for inhibitors belonging to 10 structurally different families. Using deviation in Cα-Cα distances as a criterion for loop similarity, we found that the P3-P3′ segment defines most properly the length of the loop. When conformational differences among loops of individual inhibitors were compared using root mean square deviation (rmsd) in atomic coordinates for all main chain atoms (Δr method) and rmsd operating in main chain torsion angles (Δt method), differences of up to 2.1 Å and 72.3°, respectively, were observed. Such large values indicate significant conformational differences among individual loops. Nevertheless, the overall geometry of the inhibitor-proteinase interaction is very well preserved, as judged from the similarity of Cα-Cα distances between Cα of catalytic Ser and Cα of P3-P3′ residues in various enzyme-inhibitor complexes. The mode of interaction is very well preserved both in the chymotrypsin and subtilisin families, as distances calculated for subtilisin-inhibitor complexes are almost always within the range of those for chymotrypsin-inhibitor complexes. Complex formation leads to conformational changes of up to 160° for χ1 angle. Side chains of residue P1 and P2′ adopt in different complexes a similar orientation (χ1 angle = −60° and −180°, respectively). To check whether the canonical conformation can be found among non–proteinase-inhibitor Brookhaven Protein Data Bank structures, two selection criteria—the allowed main chain dihedral angles and Cα-Cα distances for the P3-P3′ segment—were applied to all these structures. This procedure detected 132 unique hexapeptide segments in 121 structurally and functionally unrelated proteins. Partial preferences for certain amino acids occurring at particular positions in these hexapeptides could be noted. Further restriction of this set to hexapeptides with a highly exposed P1 residue side chain resulted in 40 segments. The possibility of complexes formation between these segments and serine proteinases was ruled out in molecular modeling due to steric clashes. Several structural features that determine the canonical conformation of the loop both in inhibitors and in other proteins can be distinguished. They include main chain hydrogen bonds both within the P3-P3′ segment and with the scaffold region, P3-P4 and P3′-P4′ hydrophobic interactions, and finally either hydrophobic or polar interactions involving the P1′ residue. Proteins 32:459–474, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

14.
Polcalcins are small EF‐hand proteins believed to assist in regulating pollen‐tube growth. Phl p 7, from timothy grass (Phleum pratense), crystallizes as a domain‐swapped dimer at low pH. This study describes the solution structures of the recombinant protein in buffered saline at pH 6.0, containing either 5.0 mM EDTA, 5.0 mM Mg2+, or 100 μM Ca2+. Phl p 7 is monomeric in all three ligation states. In the apo‐form, both EF‐hand motifs reside in the closed conformation, with roughly antiparallel N‐ and C‐terminal helical segments. In 5.0 mM Mg2+, the divalent ion is bound by EF‐hand 2, perturbing interhelical angles and imposing more regular helical structure. The structure of Ca2+‐bound Phl p 7 resembles that previously reported for Bet v 4—likewise exposing apolar surface to the solvent. Occluded in the apo‐ and Mg2+‐bound forms, this surface presumably provides the docking site for Phl p 7 targets. Unlike Bet v 4, EF‐hand 2 in Phl p 7 includes five potential anionic ligands, due to replacement of the consensus serine residue at –x (residue 55 in Phl p 7) with aspartate. In the Phl p 7 crystal structure, D55 functions as a helix cap for helix D. In solution, however, D55 apparently serves as a ligand to the bound Ca2+. When Mg2+ resides in site 2, the D55 carboxylate withdraws to a distance consistent with a role as an outer‐sphere ligand. 15N relaxation data, collected at 600 MHz, indicate that backbone mobility is limited in all three ligation states. Proteins 2013. © 2012 Wiley Periodicals, Inc.  相似文献   

15.
We computationally designed a de novo protein–protein interaction between wild‐type ubiquitin and a redesigned scaffold. Our strategy was to incorporate zinc at the designed interface to promote affinity and orientation specificity. A large set of monomeric scaffold surfaces were computationally engineered with three‐residue zinc coordination sites, and the ubiquitin residue H68 was docked to the open coordination site to complete a tetrahedral zinc site. This single coordination bond was intended as a hotspot and polar interaction for ubiquitin binding, and surrounding residues on the scaffold were optimized primarily as hydrophobic residues using a rotamer‐based sequence design protocol in Rosetta. From thousands of independent design simulations, four sequences were selected for experimental characterization. The best performing design, called Spelter, binds tightly to zinc (Kd < 10 nM) and binds ubiquitin with a Kd of 20 µM in the presence of zinc and 68 µM in the absence of zinc. Mutagenesis studies and nuclear magnetic resonance chemical shift perturbation experiments indicate that Spelter interacts with H68 and the target surface on ubiquitin; however, H68 does not form a hotspot as intended. Instead, mutation of H68 to alanine results in tighter binding. Although a 3/1 zinc coordination arrangement at an interface cannot be ruled out as a means to improve affinity, our study led us to conclude that 2/2 coordination arrangements or multiple‐zinc designs are more likely to promote high‐affinity protein interactions. Proteins 2013; 81:1245–1255. © 2013 Wiley Periodicals, Inc.  相似文献   

16.
An immunomodulating tetrapeptide, rigin (H‐Gly‐Gln‐Pro‐Arg‐OH), has been examined for its secondary structure preferences through combined use of high‐temperature unrestrained MD simulations in implicit water and 1D and 2D 1H NMR spectroscopy. The distribution of backbone torsion angles revealed the predominance of trans conformers across the Xaa‐Pro peptide bond. The results of MD simulations revealed that of the five predicted families A–E, the predominant families, family A (92 structures), family C (63 structures) and family D (31 structures), could be complemented by extensive 1D and 2D 1H NMR parameters acquired in aqueous PBS solution. A survey of specific inter‐ and intraresidue NOEs substantiated the predominance of an unusual type VII β‐turn structure, defined by two torsion angles, i.e. ψGln ~ 155° and ?Pro ~ ? 65° across the Gln‐Pro segment. The proposed semi‐folded kinked topology precluded formation of any intramolecular interaction, i.e. hydrogen bond or electrostatic interaction. Far‐UV CD spectral characteristics of rigin in aqueous PBS solution and non‐aqueous structure‐promoting organic solvents, TFE and TMP, revealed its strong solvent dependence. However, in aqueous PBS solution, the presence of a weak negative shoulder at nm could be ascribed to a small population with ordered, semi‐folded topology. We propose that the plausible structural attributes may be exploited for design and rigidification of the bioactive conformation of this immunomodulator toward improved immunopharmacological properties. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

17.
Kikuchi T 《Amino acids》2008,35(3):541-549
It is well-known that the IgG-binding domain from staphylococcal protein A folds into a 3α helix bundle structure, while the IgG-binding domain of streptococcal protein G forms an (α + β) structure. Recently, He et al. (Biochemistry 44:14055–14061, 2005) made mutants of these proteins from the wild types of protein A and protein G strains. These mutants are referred to as protein A219 and protein G311, and it was showed that these two mutants have different 3D structures, i.e., the 3α helix bundle structure and the (α + β) structure, respectively, despite the high sequence identity (59%). The purpose of our study was to clarify how such 3D structural differences are coded in the sequences with high homology. To address this problem, we introduce a predicted contact map constructed based on the interresidue average-distance statistics for prediction of folding properties of a protein. We refer to this map as an average distance map (ADM). Furthermore, the statistics of interresidue distances can be converted to an effective interresidue potential. We calculated the contact frequency of each residue of a protein in random conformations with this effective interresidue potential, and then we obtained values similar to ϕ values. We refer to this contact frequency of each residue as a p(μ) value. The comparison of the p(μ) values to the ϕ values for a protein suggests that p(μ) values reveal the information on the folding initiation site. Using these techniques, we try to extract the information on the difference in the 3D structures of protein A219 and protein G311 coded in their amino acid sequences in the present work. The results show that the ADM analyses and the p(μ) value analyses predict the information of folding initiation sites, which can be used to detect the 3D difference in both proteins.  相似文献   

18.
Modeling mutations in protein structures   总被引:2,自引:0,他引:2  
We describe an automated method for the modeling of point mutations in protein structures. The protein is represented by all non-hydrogen atoms. The scoring function consists of several types of physical potential energy terms and homology-derived restraints. The optimization method implements a combination of conjugate gradient minimization and molecular dynamics with simulated annealing. The testing set consists of 717 pairs of known protein structures differing by a single mutation. Twelve variations of the scoring function were tested in three different environments of the mutated residue. The best-performing protocol optimizes all the atoms of the mutated residue, with respect to a scoring function that includes molecular mechanics energy terms for bond distances, angles, dihedral angles, peptide bond planarity, and non-bonded atomic contacts represented by Lennard-Jones potential, dihedral angle restraints derived from the aligned homologous structure, and a statistical potential for non-bonded atomic interactions extracted from a large set of known protein structures. The current method compares favorably with other tested approaches, especially when predicting long and flexible side-chains. In addition to the thoroughness of the conformational search, sampled degrees of freedom, and the scoring function type, the accuracy of the method was also evaluated as a function of the flexibility of the mutated side-chain, the relative volume change of the mutated residue, and its residue type. The results suggest that further improvement is likely to be achieved by concentrating on the improvement of the scoring function, in addition to or instead of increasing the variety of sampled conformations.  相似文献   

19.
The structures of N-glycans of total glycoproteins in royal jelly have been explored to clarify whether antigenic N-glycans occur in the famous health food. The structural feature of N-glycans linked to glycoproteins in royal jelly was first characterized by immunoblotting with an antiserum against plant complex type N-glycan and lectin-blotting with Con A and WGA. For the detail structural analysis of such N-glycans, the pyridylaminated (PA-) N-glycans were prepared from hydrazinolysates of total glycoproteins in royal jelly and each PA-sugar chain was purified by reverse-phase HPLC and size-fractionation HPLC. Each structure of the PA-sugar chains purified was identified by the combination of two-dimensional PA-sugar chain mapping, ESI-MS and MS/MS analyses, sequential exoglycosidase digestions, and 500 MHz 1H-NMR spectrometry.

The immunoblotting and lectinblotting analyses preliminarily suggested the absence of antigenic N-glycan bearing β1-2 xylosyl and/or α1-3 fucosyl residue(s) and occurrence of β1-4GlcNAc residue in the insect glycoproteins.

The detailed structural analysis of N-glycans of total royal jelly glycoproteins revealed that the antigenic N-glycans do not occur but the typical high mannose-type structure (Man9~4GlcNAc2) occupies 71.6% of total N-glycan, biantennary-type structures (GlcNAc2Man3GlcNAc2) 8.4%, and hybrid type structure (GlcNAc1Man4GlcNAc2) 3.0%. Although the complete structures of the remaining 17% N-glycans; C4, (HexNAc3Hex3HexNAc2: 3.0%), D2 (HexNAc2Hex5HexNAc2: 4.5%), and D3 (HexNAc3Hex4HexNAc2: 9.5%) are still obscure so far, ESI-MS analysis, exoglycosidase digestions by two kinds of β-N-acetylglucosaminidase, and WGA blotting suggested that these N-glycans might bear a β1-4 linkage N-acetylglucosaminyl residue.  相似文献   

20.
In the absence of experimentally determined protein structure many biological questions can be addressed using computational structural models. However, the utility of protein structural models depends on their quality. Therefore, the estimation of the quality of predicted structures is an important problem. One of the approaches to this problem is the use of knowledge‐based statistical potentials. Such methods typically rely on the statistics of distances and angles of residue‐residue or atom‐atom interactions collected from experimentally determined structures. Here, we present VoroMQA (Voronoi tessellation‐based Model Quality Assessment), a new method for the estimation of protein structure quality. Our method combines the idea of statistical potentials with the use of interatomic contact areas instead of distances. Contact areas, derived using Voronoi tessellation of protein structure, are used to describe and seamlessly integrate both explicit interactions between protein atoms and implicit interactions of protein atoms with solvent. VoroMQA produces scores at atomic, residue, and global levels, all in the fixed range from 0 to 1. The method was tested on the CASP data and compared to several other single‐model quality assessment methods. VoroMQA showed strong performance in the recognition of the native structure and in the structural model selection tests, thus demonstrating the efficacy of interatomic contact areas in estimating protein structure quality. The software implementation of VoroMQA is freely available as a standalone application and as a web server at http://bioinformatics.lt/software/voromqa . Proteins 2017; 85:1131–1145. © 2017 Wiley Periodicals, Inc.  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