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.
The reversible acetylation of lysine to form N6‐acetyllysine in the regulation of protein function is a hallmark of epigenetics. Acetylation of the positively charged amino group of the lysine side chain generates a neutral N‐alkylacetamide moiety that serves as a molecular “switch” for the modulation of protein function and protein–protein interactions. We now report the analysis of 381 N6‐acetyllysine side chain amide conformations as found in 79 protein crystal structures and 11 protein NMR structures deposited in the Protein Data Bank (PDB) of the Research Collaboratory for Structural Bioinformatics. We find that only 74.3% of N6‐acetyllysine residues in protein crystal structures and 46.5% in protein NMR structures contain amide groups with energetically preferred trans or generously trans conformations. Surprisingly, 17.6% of N6‐acetyllysine residues in protein crystal structures and 5.3% in protein NMR structures contain amide groups with energetically unfavorable cis or generously cis conformations. Even more surprisingly, 8.1% of N6‐acetyllysine residues in protein crystal structures and 48.2% in NMR structures contain amide groups with energetically prohibitive twisted conformations that approach the transition state structure for cistrans isomerization. In contrast, 109 unique N‐alkylacetamide groups contained in 84 highly accurate small molecule crystal structures retrieved from the Cambridge Structural Database exclusively adopt energetically preferred trans conformations. Therefore, we conclude that cis and twisted N6‐acetyllysine amides in protein structures deposited in the PDB are erroneously modeled due to their energetically unfavorable or prohibitive conformations. Proteins 2013; © 2012 Wiley Periodicals, Inc.  相似文献   

2.
High‐resolution homology models are useful in structure‐based protein engineering applications, especially when a crystallographic structure is unavailable. Here, we report the development and implementation of RosettaAntibody, a protocol for homology modeling of antibody variable regions. The protocol combines comparative modeling of canonical complementarity determining region (CDR) loop conformations and de novo loop modeling of CDR H3 conformation with simultaneous optimization of VL‐VH rigid‐body orientation and CDR backbone and side‐chain conformations. The protocol was tested on a benchmark of 54 antibody crystal structures. The median root mean square deviation (rmsd) of the antigen binding pocket comprised of all the CDR residues was 1.5 Å with 80% of the targets having an rmsd lower than 2.0 Å. The median backbone heavy atom global rmsd of the CDR H3 loop prediction was 1.6, 1.9, 2.4, 3.1, and 6.0 Å for very short (4–6 residues), short (7–9), medium (10–11), long (12–14) and very long (17–22) loops, respectively. When the set of ten top‐scoring antibody homology models are used in local ensemble docking to antigen, a moderate‐to‐high accuracy docking prediction was achieved in seven of fifteen targets. This success in computational docking with high‐resolution homology models is encouraging, but challenges still remain in modeling antibody structures for sequences with long H3 loops. This first large‐scale antibody–antigen docking study using homology models reveals the level of “functional accuracy” of these structural models toward protein engineering applications. Proteins 2009; 74:497–514. © 2008 Wiley‐Liss, Inc.  相似文献   

3.
A blinded study to assess the state of the art in three‐dimensional structure modeling of the variable region (Fv) of antibodies was conducted. Nine unpublished high‐resolution x‐ray Fab crystal structures covering a wide range of antigen‐binding site conformations were used as benchmark to compare Fv models generated by four structure prediction methodologies. The methodologies included two homology modeling strategies independently developed by CCG (Chemical Computer Group) and Accerlys Inc, and two fully automated antibody modeling servers: PIGS (Prediction of ImmunoGlobulin Structure), based on the canonical structure model, and Rosetta Antibody Modeling, based on homology modeling and Rosetta structure prediction methodology. The benchmark structure sequences were submitted to Accelrys and CCG and a set of models for each of the nine antibody structures were generated. PIGS and Rosetta models were obtained using the default parameters of the servers. In most cases, we found good agreement between the models and x‐ray structures. The average rmsd (root mean square deviation) values calculated over the backbone atoms between the models and structures were fairly consistent, around 1.2 Å. Average rmsd values of the framework and hypervariable loops with canonical structures (L1, L2, L3, H1, and H2) were close to 1.0 Å. H3 prediction yielded rmsd values around 3.0 Å for most of the models. Quality assessment of the models and the relative strengths and weaknesses of the methods are discussed. We hope this initiative will serve as a model of scientific partnership and look forward to future antibody modeling assessments. Proteins 2011; © 2011 Wiley‐Liss, Inc.  相似文献   

4.
Lee HS  Zhang Y 《Proteins》2012,80(1):93-110
We developed BSP‐SLIM, a new method for ligand–protein blind docking using low‐resolution protein structures. For a given sequence, protein structures are first predicted by I‐TASSER; putative ligand binding sites are transferred from holo‐template structures which are analogous to the I‐TASSER models; ligand–protein docking conformations are then constructed by shape and chemical match of ligand with the negative image of binding pockets. BSP‐SLIM was tested on 71 ligand–protein complexes from the Astex diverse set where the protein structures were predicted by I‐TASSER with an average RMSD 2.92 Å on the binding residues. Using I‐TASSER models, the median ligand RMSD of BSP‐SLIM docking is 3.99 Å which is 5.94 Å lower than that by AutoDock; the median binding‐site error by BSP‐SLIM is 1.77 Å which is 6.23 Å lower than that by AutoDock and 3.43 Å lower than that by LIGSITECSC. Compared to the models using crystal protein structures, the median ligand RMSD by BSP‐SLIM using I‐TASSER models increases by 0.87 Å, while that by AutoDock increases by 8.41 Å; the median binding‐site error by BSP‐SLIM increase by 0.69Å while that by AutoDock and LIGSITECSC increases by 7.31 Å and 1.41 Å, respectively. As case studies, BSP‐SLIM was used in virtual screening for six target proteins, which prioritized actives of 25% and 50% in the top 9.2% and 17% of the library on average, respectively. These results demonstrate the usefulness of the template‐based coarse‐grained algorithms in the low‐resolution ligand–protein docking and drug‐screening. An on‐line BSP‐SLIM server is freely available at http://zhanglab.ccmb.med.umich.edu/BSP‐SLIM . Proteins 2012. © 2011 Wiley Periodicals, Inc.  相似文献   

5.
While the cis‐acyltransferase modular polyketide synthase assembly lines have largely been structurally dissected, enzymes from within the recently discovered trans‐acyltransferase polyketide synthase assembly lines are just starting to be observed crystallographically. Here we examine the ketoreductase (KR) from the first polyketide synthase module of the bacillaene nonribosomal peptide synthetase/polyketide synthase at 2.35‐Å resolution. This KR naturally reduces both α‐ and β‐keto groups and is the only KR known to do so during the biosynthesis of a polyketide. The isolated KR not only reduced an N‐acetylcysteamine‐bound β‐keto substrate to a D ‐β‐hydroxy product, but also an N‐acetylcysteamine‐bound α‐keto substrate to an L ‐α‐hydroxy product. That the substrates must enter the active site from opposite directions to generate these stereochemistries suggests that the acyl‐phosphopantetheine moiety is capable of accessing very different conformations despite being anchored to a serine residue of a docked acyl carrier protein. The features enabling stereocontrolled α‐ketoreduction may not be extensive since a KR that naturally reduces a β‐keto group within a cis‐acyltransferase polyketide synthase was identified that performs a completely stereoselective reduction of the same α‐keto substrate to generate the D ‐α‐hydroxy product. A sequence analysis of trans‐acyltransferase KRs reveals that a single residue, rather than a three‐residue motif found in cis‐acyltransferase KRs, is predictive of the orientation of the resulting β‐hydroxyl group. Proteins 2014; 82:2067–2077. © 2014 Wiley Periodicals, Inc.  相似文献   

6.
The crystal structure of the GH78 family α‐rhamnosidase from Klebsiella oxytoca (KoRha) has been determined at 2.7 Å resolution with rhamnose bound in the active site of the catalytic domain. Curiously, the putative catalytic acid, Asp 222, is preceded by an unusual non‐proline cis‐peptide bond which helps to project the carboxyl group into the active centre. This KoRha homodimeric structure is significantly smaller than those of the other previously determined GH78 structures. Nevertheless, the enzyme displays α‐rhamnosidase activity when assayed in vitro, suggesting that the additional structural domains found in the related enzymes are dispensible for function. Proteins 2015; 83:1742–1749. © 2015 The Authors. Proteins: Structure, Function, and Bioinformatics Published by Wiley Periodicals, Inc.  相似文献   

7.
To investigate the role of proline in defining β turn conformations within cyclic hexa- and pentapeptides we synthesized and determined the conformations of a series of L - and D -proline-containing peptides by means of 2D NMR spectroscopy and restrained molecular dynamics simulations. Due to cis/trans isomerism the L -proline peptides adopt at least two different conformations that are analyzed and compared to the structures of the corresponding D -proline peptides. The cis conformations of the compounds cyclo(-Pro-Ala-Ala-Pro-Ala-Ala-), cyclo(-Arg-Gly-Asp-Phe-Pro-Gly-), cyclo(-Arg-Gly-Asp-Phe-Pro-Ala-), cyclo(-Pro-Ala-Ala-Ala-Ala--), and cyclo(-Pro-Ala-Pro-Ala-Ala-) form uncommon βVI turns that mimic the turn geometries found in crystallographically refined protein structures at such a detailed level that even preferred side chain orientations are reproduced. The ratios of the cis/trans isomers are analyzed in terms of the steric demand of the proline-following residue. The conformational details derived from this study illustrate the importance of the examination of small model compounds derived from protein loop regions, especially if bioactive recognition sequences, such as RGD (Arg-Gly-Asp), are incorporated. © 1993 Wiley-Liss, Inc.  相似文献   

8.
A conserved cis proline residue located in the active site of Thermotoga maritima acetyl esterase (TmAcE) from the carbohydrate esterase family 7 (CE7) has been substituted by alanine. The residue was known to play a crucial role in determining the catalytic properties of the enzyme. To elucidate the structural role of the residue, the crystal structure of the Pro228Ala variant (TmAcEP228A) was determined at 2.1 Å resolution. The replacement does not affect the overall secondary, tertiary, and quaternary structures and moderately decreases the thermal stability. However, the wild type cis conformation of the 227–228 peptide bond adopts a trans conformation in the variant. Other conformational changes in the tertiary structure are restricted to residues 222–226, preceding this peptide bond and are located away from the active site. Overall, the results suggest that the conserved proline residue is responsible for the cis conformation of the peptide and shapes the geometry of the active site. Elimination of the pyrrolidine ring results in the loss of van der Waals and hydrophobic interactions with both the alcohol and acyl moeities of the ester substrate, leading to significant impairment of the activity and perturbation of substrate specificity. Furthermore, a cis‐to‐trans conformational change arising out of residue changes at this position may be associated with the evolution of divergent activity, specificity, and stability properties of members constituting the CE7 family. Proteins 2017; 85:694–708. © 2016 Wiley Periodicals, Inc.  相似文献   

9.
The endogenous Escherichia coli porin OmpF was crystallized as an accidental by‐product of our efforts to express, purify, and crystallize the E. coli integral membrane protein KdpD in the presence of foscholine‐12 (FC12). FC12 is widely used in membrane protein studies, but no crystal structure of a protein that was both purified and crystallized with this detergent has been reported in the Protein Data Bank. Crystallization screening for KdpD yielded two different crystals of contaminating protein OmpF. Here, we report two OmpF structures, the first membrane protein crystal structures for which extraction, purification, and crystallization were done exclusively with FC12. The first structure was refined in space group P21 with cell parameters a = 136.7 Å, b = 210.5 Å, c = 137 Å, and β = 100.5°, and the resolution of 3.8 Å. The second structure was solved at the resolution of 4.4 Å and was refined in the P321 space group, with unit cell parameters a = 215.5 Å, b = 215.5 Å, c = 137.5 Å, and γ = 120°. Both crystal forms show novel crystal packing, in which the building block is a tetrahedral arrangement of four trimers. Additionally, we discuss the use of FC12 for membrane protein crystallization and structure determination, as well as the problem of the OmpF contamination for membrane proteins overexpressed in E. coli.  相似文献   

10.
The mechanosensitive channel of small conductance (MscS) contributes to the survival of bacteria during osmotic downshock by transiently opening large diameter pores for the efflux of cellular contents before the membrane ruptures. Two crystal structures of the Escherichia coli MscS are currently available, the wild type protein in a nonconducting state at 3.7 Å resolution (Bass et al., Science 2002; 298:1582–1587) and the Ala106Val variant in an open state at 3.45 Å resolution (Wang et al., Science 2008; 321:1179–1183). Both structures used protein solubilized in the detergent fos‐choline‐14. We report here crystal structures of MscS from E. coli and Helicobacter pylori solubilized in the detergent β‐dodecylmaltoside at resolutions of 4.4 and 4.2 Å, respectively. While the cytoplasmic domains are unchanged in these structures, distinct conformations of the transmembrane domains are observed. Intriguingly, β‐dodecylmaltoside solubilized wild type E. coli MscS adopts the open state structure of A106V E. coli MscS, while H. pylori MscS resembles the nonconducting state structure observed for fos‐choline‐14 solubilized E. coli MscS. These results highlight the sensitivity of membrane protein conformational equilibria to variations in detergent, crystallization conditions, and protein sequence.  相似文献   

11.
Predicting the conformations of loops is a critical aspect of protein comparative (homology) modeling. Despite considerable advances in developing loop prediction algorithms, refining loops in homology models remains challenging. In this work, we use antibodies as a model system to investigate strategies for more robustly predicting loop conformations when the protein model contains errors in the conformations of side chains and protein backbone surrounding the loop in question. Specifically, our test system consists of partial models of antibodies in which the “scaffold” (i.e., the portion other than the complementarity determining region, CDR, loops) retains native backbone conformation, whereas the CDR loops are predicted using a combination of knowledge‐based modeling (H1, H2, L1, L2, and L3) and ab initio loop prediction (H3). H3 is the most variable of the CDRs. Using a previously published method, a test set of 10 shorter H3 loops (5–7 residues) are predicted to an average backbone (N? Cα? C? O) RMSD of 2.7 Å while 11 longer loops (8–9 residues) are predicted to 5.1 Å, thus recapitulating the difficulties in refining loops in models. By contrast, in control calculations predicting the same loops in crystal structures, the same method reconstructs the loops to an average of 0.5 and 1.4 Å for the shorter and longer loops, respectively. We modify the loop prediction method to improve the ability to sample near‐native loop conformations in the models, primarily by reducing the sensitivity of the sampling to the loop surroundings, and allowing the other CDR loops to optimize with the H3 loop. The new method improves the average accuracy significantly to 1.3 Å RMSD and 3.1 Å RMSD for the shorter and longer loops, respectively. Finally, we present results predicting 8–10 residue loops within complete comparative models of five nonantibody proteins. While anecdotal, these mixed, full‐model results suggest our approach is a promising step toward more accurately predicting loops in homology models. Furthermore, while significant challenges remain, our method is a potentially useful tool for predicting antibody structures based on a known Fv scaffold. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

12.
Vanillyl-alcohol oxidase catalyses the oxidation of several 4-hydroxybenzyl alcohols by using 8-α-(N3-histidyl)-FAD as a covalently bound prosthetic group. Crystals of vanillyl-alcohol oxidase from Penicillium simplicissimum have been grown by using the vapor diffusion technique. The space group was found to be I4, with cell dimensions a = b = 140.5 Å, c = 132.9 Å. Diffraction data have been recorded to 3.2 Å resolution by using a laboratory source and to 2.5 Å resolution on flash freezing the crystal at the ELETTRA Synchrotron X-ray diffraction beam line. Proteins 27:601–603, 1997. © 1997 Wiley-Liss Inc.  相似文献   

13.
The wild type red fluorescent protein eqFP578 (from sea anemone Entacmaea quadricolor, λex = 552 nm, λem = 578 nm) and its bright far‐red fluorescent variant Katushka (λex = 588 nm, λem = 635 nm) are characterized by the pronounced pH dependence of their fluorescence. The crystal structures of eqFP578f (eqFP578 with two point mutations improving the protein folding) and Katushka have been determined at the resolution ranging from 1.15 to 1.85 Å at two pH values, corresponding to low and high level of fluorescence. The observed extinguishing of fluorescence upon reducing pH in eqFP578f and Katushka has been shown to be accompanied by the opposite transcis and cis‐trans chromophore isomerization, respectively. Asn143, Ser158, His197 and Ser143, Leu174, and Arg197 have been shown to stabilize the respective trans and cis fluorescent states of the chromophores in eqFP578f and Katushka at higher pH. The cis state has been suggested as being primarily responsible for the observed far‐red shift of the emission maximum of Katushka relative to that of eqFP578f.  相似文献   

14.
A particulate membrane fraction from Saccharomyces cerevisiae contains transferases which catalyze the incorporation of N-acetylglucosamine from UDP-N-acetylglucosamine into a lipid fraction as well as into a protein fraction. The lipid fraction contains two alkali-stable lipids which can be separated on a silica G-60 column. The sugar moieties of these polyprenoid lipids are: N-acetylglucosamine and di-N-acetylchitobiose. The transfer of carbohydrate from isolated glycolipids to endogenous protein has been examined. After separation of protein and saccharide by hydrazinolysis and reacetylation only di-N-acetylchitobiose is found, and also when glycolipid containing only one N-acetylglucosamine is used as substrate. Maximum transfer of saccharides from glycolipids to protein is obtained at a Triton X-100 concentration of 1%. At this Triton X-100 concentration there is practically no transfer of N-acetylglucosamine from UDP-N-acetylglucosamine to the phosphorylated lipid. Therefore, when polyprenyl diphosphate N-acetyl[3H]-glucosamine is incubated together with UDP-N-acetyl[14C]glucosamine with the membrane fraction in the presence of 1% Triton X-100, a doubly labelled di-N-acetylchitobiose linked to lipid is formed with N-acetyl[14C]glucosamine at the non-reducing end of the chain.  相似文献   

15.
Mark E. Snow 《Proteins》1993,15(2):183-190
A novel scheme for the parameterization of a type of “potential energy” function for protein molecules is introduced. The function is parameterized based on the known conformations of previously determined protein structures and their sequence similarity to a molecule whose conformation is to be calculated. Once parameterized, minima of the potential energy function can be located using a version of simulated annealing which has been previously shown to locate global and near-global minima with the given functional form. As a test problem, the potential was parameterized based on the known structures of the rubredoxins from Desulfovibrio vulgaris, Desulfovibrio desulfuricans, and Clostridium pasteurianum, which vary from 45 to 54 amino acids in length, and the sequence alignments of these molecules with the rubredoxin sequence from Desulfovibrio gigas. Since the Desulfovibrio gigas rubredeoxin conformation has also been determined, it is possible to check the accuracy of the results. Ten simulated-annealing runs from random starting conformations were performed. Seven of the 10 resultant conformations have an all-Cα rms deviation from the crystallographically determined conformation of less than 1.7 Å. For five of the structures, the rms deviation is less than 0.8 Å. Four of the structures have conformations which are virtually identical to each other except for the position of the carboxy-terminal residue. This is also the conformation which is achieved if the determined crystal structure is minimized with the same potential. The all-Cα rms difference between the crystal and minimized crystal structures is 0.6 Å. It is further observed that the “energies” of the structures according to the potential function exhibit a strong correlation with rms deviation from the native structure. The conformations of the individual model structures and the computational aspects of the modeling procedure are discussed. © 1993 Wiley-Liss, Inc.  相似文献   

16.
The recent technological advances in electron microscopes, detectors, as well as image processing and reconstruction software have brought single particle cryo-electron microscopy (cryo-EM) into prominence for determining structures of bio-molecules at near atomic resolution. This has been particularly true for virus capsids, ribosomes, and other large assemblies, which have been the ideal specimens for structural studies by cryo-EM approaches. An analysis of time series metadata of virus structures on the methods of structure determination, resolution of the structures, and size of the virus particles revealed a rapid increase in the virus structures determined by cryo-EM at near atomic resolution since 2010. In addition, the data highlight the median resolution (~3.0?Å) and size (~310.0?Å in diameter) of the virus particles determined by X-ray crystallography while no such limits exist for cryo-EM structures, which have a median diameter of 508?Å. Notably, cryo-EM virus structures in the last four years have a median resolution of 3.9?Å. Taken together with minimal sample requirements, not needing diffraction quality crystals, and being able to achieve similar resolutions of the crystal structures makes cryo-EM the method of choice for current and future virus capsid structure determinations.  相似文献   

17.
Eight conformers typical of diketopiperazine (DKP) ring folding were chosen for analysis. Conformational energy calculations were carried out using the semiempirical quantummechanical CNDO/2 method. The results obtained confirm considerable flexibility of the DKP skeleton. As the degree of folding increases, twisted boat conformations with the nonplanar peptide bonds tend to be more stable, while more rigid regular boat conformations with planar peptide bonds appear to be less stable than a flat one. The CNDO/2 method was found to be reliable enough for conformational studies of cyclic peptide skeletons with cis-peptide bonds.  相似文献   

18.
Semi-empirical energy calculations for an internal Pro-Pro dimer are presented that take into account the nature of the flexibility of the proline ring due to its puckering. Calculations show that three stable conformations are available for the dimer: the cis (ω = 0°, ψ = 160°); the trans (ω = 180°, ψ = 160°, also referred to as trans′); and the cis′ (ω = 180°, ψ = ?40°) conformations. The best conformational pathways between these stable conformations are determined. Calculations also show that the barrier for cis′–trans′ conversion is of the same order of magnitude as that for cistrans conversion.  相似文献   

19.
In this study, we report two high‐resolution structures of the pyridoxal 5′ phosphate (PLP)‐dependent enzyme kynurenine aminotransferase‐I (KAT‐I). One is the native structure with the cofactor in the PLP form bound to Lys247 with the highest resolution yet available for KAT‐I at 1.28 Å resolution, and the other with the general PLP‐dependent aminotransferase inhibitor, aminooxyacetate (AOAA) covalently bound to the cofactor at 1.54 Å. Only small conformational differences are observed in the vicinity of the aldimine (oxime) linkage with which the PLP forms the Schiff base with Lys247 in the 1.28 Å resolution native structure, in comparison to other native PLP‐bound structures. We also report the inhibition of KAT‐1 by AOAA and aminooxy‐phenylpropionic acid (AOPP), with IC50s of 13.1 and 5.7 μM, respectively. The crystal structure of the enzyme in complex with the inhibitor AOAA revealed that the cofactor is the PLP form with the external aldimine linkage. The location of this oxime with the PLP, which forms in place of the native internal aldimine linkage of PLP of the native KAT‐I, is away from the position of the native internal aldimine, with the free Lys247 substantially retaining the orientation of the native structure. Tyr101, at the active site, was observed in two conformations in both structures.  相似文献   

20.
Enzymes of the glyoxylate shunt are important for the virulence of pathogenic organisms such as Mycobacterium tuberculosis and Candida albicans. Two isoforms have been identified for malate synthase, the second enzyme in the pathway. Isoform A, found in fungi and plants, comprises ~530 residues, whereas isoform G, found only in bacteria, is larger by ~200 residues. Crystal structures of malate synthase isoform G from Escherichia coli and Mycobacterium tuberculosis were previously determined at moderate resolution. Here we describe crystal structures of E. coli malate synthase A (MSA) in the apo form (1.04 Å resolution) and in complex with acetyl‐coenzyme A and a competitive inhibitor, possibly pyruvate or oxalate (1.40 Å resolution). In addition, a crystal structure for Bacillus anthracis MSA at 1.70 Å resolution is reported. The increase in size between isoforms A and G can be attributed primarily to an inserted α/β domain that may have regulatory function. Upon binding of inhibitor or substrate, several active site loops in MSA undergo large conformational changes. However, in the substrate bound form, the active sites of isoforms A and G from E. coli are nearly identical. Considering that inhibitors bind with very similar affinities to both isoforms, MSA is as an excellent platform for high‐resolution structural studies and drug discovery efforts.  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