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.
Di- and oligopeptide- binding protein OppAs play important roles in solute and nutrient uptake, sporulation, biofilm formation, cell wall muropeptides recycling, peptide-dependent quorum-sensing responses, adherence to host cells, and a variety of other biological processes. Soluble OppA from Thermoanaerobacter tengcongensis was expressed in Escherichia coli. The protein was found to be >95% pure with SDS–PAGE after a series of purification steps and the purity was further verified by mass spectrometry. The protein was crystallized using the sitting-drop vapour-diffusion method with PEG 400 as the precipitant. Crystal diffraction extended to 2.25 Å. The crystal belonged to space group C2221, with unit-cell parameters of a = 69.395, b = 199.572, c = 131.673 Å, and α = β = γ = 90°.  相似文献   

2.
Human astroviruses (HAstVs) are a major cause of gastroenteritis. HAstV assembles from the structural protein VP90 and undergoes a cascade of proteolytic cleavages. Cleavage to VP70 is required for release of immature particles from cells, and subsequent cleavage by trypsin confers infectivity. We used electron cryomicroscopy and icosahedral image analysis to determine the first experimentally derived, three-dimensional structures of an immature VP70 virion and a fully proteolyzed, infectious virion. Both particles display T = 3 icosahedral symmetry and nearly identical solid capsid shells with diameters of ~ 350 Å. Globular spikes emanate from the capsid surface, yielding an overall diameter of ~ 440 Å. While the immature particles display 90 dimeric spikes, the mature capsid only displays 30 spikes, located on the icosahedral 2-fold axes. Loss of the 60 peripentonal spikes likely plays an important role in viral infectivity. In addition, immature HAstV bears a striking resemblance to the structure of hepatitis E virus (HEV)-like particles, as previously predicted from structural similarity of the crystal structure of the astrovirus spike domain with the HEV P-domain [Dong, J., Dong, L., Méndez, E. &; Tao, Y. (2011). Crystal structure of the human astrovirus capsid spike. Proc. Natl. Acad. Sci. USA 108, 12681–12686]. Similarities between their capsid shells and dimeric spikes and between the sequences of their capsid proteins suggest that these viral families are phylogenetically related and may share common assembly and activation mechanisms.  相似文献   

3.
Recent advances in three-dimensional electron microscopy (3D EM) have enabled the quantitative visualization of the structural building blocks of proteins at improved resolutions. We provide algorithms to detect the secondary structures (α-helices and β-sheets) from proteins for which the volumetric maps are reconstructed at 6–10 Å resolution. Additionally, we show that when the resolution is coarser than 10 Å, some of the supersecondary structures can be detected from 3D EM maps. For both these algorithms, we employ tools from computational geometry and differential topology, specifically the computation of stable/unstable manifolds of certain critical points of the distance function induced by the molecular surface. Our results connect mathematically well-defined constructions with bio-chemically induced structures observed in proteins.  相似文献   

4.
Dietary nucleotides have been shown to benefit many physiological and nutritional functions in higher vertebrates and fish. Therefore, a 6-week feeding trial was conducted to evaluate the effects of graded levels of a commercial nucleotide product on growth performance, immune responses and intestinal morphology of juvenile red drum (initial average weight of 7.1 g). The basal diet was formulated to contain 40% protein, 10% lipid and a digestible energy level of 3.5 kcal g?1. Two levels of nucleotide (Ascogen P®, 0.5% and 1% of diet) were added to the basal diet with menhaden fishmeal and menhaden oil adjusted to provide isonitrogenous and isolipidic diets. Nucleotide supplementation tended to improve weight gain and survival of red drum, but not at a significant level. Neutrophil oxidative radical anion production and serum lysozyme activity tended to be higher for fish fed diets supplemented with nucleotide, while extracellular superoxide anion production of head kidney macrophages from fish fed diets with 1% nucleotide was significantly (P < 0.05) increased, although no significant differences were observed between fish fed 0.5% nucleotide diet and the basal diet.Nucleotide supplementation significantly (P < 0.05) increased fold height in the proximal intestine, and enterocyte height in the pyloric caeca, proximal and distal enteric sections. A significantly (P < 0.05) higher microvilli height was observed in all evaluated enteric sections of fish fed with diets supplemented with nucleotides. It is therefore possible to use dietary nucleotides supplementation to significantly enhance the intestinal structure of red drum. Likewise, nucleotides in the diet may improve some components of the non-specific immune response of this sciaenid fish.  相似文献   

5.
6.
Heterodimeric nucleotide binding domains NBD1/NBD2 distinguish the ATP-binding cassette protein SUR2A, a recognized regulatory subunit of cardiac ATP-sensitive K+ (KATP) channels. The tandem function of these core domains ensures metabolism-dependent gating of the Kir6.2 channel pore, yet their structural arrangement has not been resolved. Here, purified monodisperse and interference-free recombinant particles were subjected to synchrotron radiation small-angle X-ray scattering (SAXS) in solution. Intensity function analysis of SAXS profiles resolved NBD1 and NBD2 as octamers. Implemented by ab initio simulated annealing, shape determination prioritized an oblong envelope wrapping NBD1 and NBD2 with respective dimensions of 168 × 80 × 37 Å3 and 175 × 81 × 37 Å3 based on symmetry constraints, validated by atomic force microscopy. Docking crystal structure homology models against SAXS data reconstructed the NBD ensemble surrounding an inner cleft suitable for Kir6.2 insertion. Human heart disease-associated mutations introduced in silico verified the criticality of the mapped protein–protein interface. The resolved quaternary structure delineates thereby a macromolecular arrangement of KATP channel SUR2A regulatory domains.  相似文献   

7.
Formation of spider silk from its constituent proteins—spidroins—involves changes from soluble helical/coil conformations to insoluble β-sheet aggregates. This conversion needs to be regulated to avoid precocious aggregation proximally in the silk gland while still allowing rapid silk assembly in the distal parts. Lowering of pH from about 7 to 6 is apparently important for silk formation. The spidroin N-terminal domain (NT) undergoes stable dimerization and structural changes in this pH region, but the underlying mechanisms are incompletely understood. Here, we determine the NMR and crystal structures of Euprosthenops australis NT mutated in the dimer interface (A72R). Also, the NMR structure of wild‐type (wt) E. australis NT at pH 7.2 and 300 mM sodium chloride was determined. The wt NT and A72R structures are monomers and virtually identical, but they differ from the subunit structure of dimeric wt NT mainly by having a tryptophan (W10) buried between helix 1 and helix 3, while W10 is surface exposed in the dimer. Wedging of the W10 side chain in monomeric NT tilts helix 3 approximately 5–6 Å into a position that is incompatible with that of the observed dimer structure. The structural differences between monomeric and dimeric NT domains explain the tryptophan fluorescence patterns of NT at pH 7 and pH 6 and indicate that the biological function of NT depends on conversion between the two conformations.  相似文献   

8.
The crystal structure of Aspergillus oryzae carbonic anhydrase (AoCA) was determined at 2.7 Å resolution and it revealed a dimer, which only has precedents in the α class in two membrane and cancer-associated enzymes. α carbonic anhydrases are underrepresented in fungi compared to the β class, this being the first structural representative. The overall fold and zinc binding site resemble other well studied carbonic anhydrases. A major difference is that the histidine, thought to be the major proton shuttle residue in most mammalian enzymes, is replaced by a phenylalanine in AoCA. This finding poses intriguing questions as to the biological functions of fungal α carbonic anhydrases, which are promising candidates for biotechnological applications.Structured summaryAoCA binds to AoCA by molecular sieving (View interaction)AoCA binds to AoCA X-ray crystallography (View interaction)  相似文献   

9.
We investigated the suitability of surface plasmon resonance (SPR) for providing quantitative binding information from direct screening of a chemical library on protein tyrosine phosphatase 1b (PTP1B). The experimental design was established from simulations to detect binding with KD < 10?4 M. The 1120 compounds (cpds) were injected sequentially at concentrations [C(cpd)] of 0.5 or 10 μM over various target surfaces. An optimized evaluation procedure was applied. More than 90% of cpds showed no detectable signal in four screens. The 30 highest responders at C(cpd) = 10 μM, of which 25 were selected in at least one of three screens at C(cpd) = 0.5 μM, contained 22 promiscuous binders and 8 potential PTP1B-specific binders with KD  10?5 M. Inhibition of PTP1B activity was assayed and confirmed for 6 of these, including sanguinarine, a known PTP1B inhibitor. C(cpd) dependence studies fully confirmed screening conclusions. The quantitative consistency of SPR data led us to propose a structure–activity relationship (SAR) model for developing selective PTP1B inhibitors based on the ranking of 10 arylbutylpiperidine analogs.  相似文献   

10.
Chen C  Kim HL  Zhuang N  Seo KH  Park KH  Han CD  Park YS  Lee KH 《FEBS letters》2011,585(17):2640-2646
Up to now, d-threo-tetrahydrobiopterin (DH4, dictyopterin) was detected only in Dictyostelium discoideum, while the isomer l-erythro-tetrahydrobioterin (BH4) is common in mammals. To elucidate the mechanism of DH4 regeneration by D. discoideum dihydropteridine reductase (DicDHPR), we have determined the crystal structure of DicDHPR complexed with NAD+ at 2.16 Å resolution. Significant structural differences from mammalian DHPRs are found around the coenzyme binding site, resulting in a higher Km value for NADH (Km = 46.51 ± 0.4 μM) than mammals. In addition, we have found that rat DHPR as well as DicDHPR could bind to both substrates quinonoid-BH2 and quinonoid-DH2 by docking calculations and have confirmed their catalytic activity by in vitro assay.Structured summary of protein interactionsDHPR binds to DHPR by X-ray crystallography (View interaction)  相似文献   

11.
Plant lectins have been studied as histological markers and promising antineoplastic molecules for a long time, and structural characterization of different lectins bound to specific cancer epitopes has been carried out successfully. The crystal structures of Vatairea macrocarpa (VML) seed lectin in complex with GalNAc-α-O-Ser (Tn antigen) and GalNAc have been determined at the resolution of 1.4 Å and 1.7 Å, respectively. Molecular docking analysis of this new structure and other Tn-binding legume lectins to O-mucin fragments differently decorated with this antigen provides a comparative binding profile among these proteins, stressing that subtle alterations that may not influence monosaccharide binding can, nonetheless, directly impact the ability of these lectins to recognize naturally occurring antigens. In addition to the specific biological effects of VML, the structural and binding similarities between it and other lectins commonly used as histological markers (e.g., VVLB4 and SBA) strongly suggest VML as a candidate tool for cancer research.  相似文献   

12.
DNA from porcine circovirus type 1 (PCV1) and 2 (PCV2) has recently been detected in two vaccines against rotaviral gastroenteritis from manufacturers A and B. We investigated if PCV1 sequences are present in other viral vaccines. We screened seeds, bulks and final vaccine preparations from ten manufacturers using qRT-PCR. We detected 3.8 × 103 to 1.9 × 107 PCV1 DNA copies/milliliter in live poliovirus seeds for inactivated polio vaccine (IPV) from manufacturer A, however, following inactivation and purification, the finished IPV was PCV1-negative. PCV1 DNA was not detectable in live polio preparations from other vaccine producers. There was no detectable PCV1 DNA in the measles, mumps, rubella and influenza vaccines analysed including material supplied by manufacturer A. We confirmed that the PCV1 genome in the rotavirus vaccine from manufacturer A is near full-length. It contains two mutations in the PCV cap gene, which may result from viral adaptation to Vero cells. Bulks of this vaccine contained 9.8 × 1010 to 1.8 × 1011 PCV1 DNA copies/millilitre and between 4.1 × 107 and 5.5 × 108 DNA copies were in the final doses. We found traces of PCV1 and PCV2 DNA in the rotavirus vaccine from manufacturer B. This highlights the issue of vaccine contamination and may impact on vaccine quality control.  相似文献   

13.
ObjectiveMyasthenia gravis (MG) is a T- and B-cell mediated autoimmune disorder affecting the neuromuscular junction. The receptor for advanced glycation endproducts (RAGE) plays a role in the amplification of chronic inflammatory disorders and autoimmune diseases. We sought to investigate the role of RAGE and its ligands in the pathophysiology of MG.MethodsIn this cross-sectional study we enrolled 42 patients with MG and 36 volunteers. We employed enzyme-linked immunosorbent assays to determine the concentration of soluble RAGE (sRAGE) and high mobility group box 1 (HMGB1) in serum of patients and volunteers. In a subpopulation of patients we measured the serum levels of endogenous secretory (es) RAGE and various RAGE ligands, such as S100B, S100A8 and advanced glycation endproducts (AGE-CML). Reported are means and standard error mean.ResultsWe found significantly reduced levels of the soluble receptors sRAGE and esRAGE in patients with MG compared to volunteers without MG (sRAGE [pg/ml] 927.2 ± 80.8 vs. 1400.1 ± 92.4; p < 0.001; esRAGE [pg/ml] 273.5 ± 24.6 vs. 449.0 ± 22.4; p < 0.001). Further categorization of patients with MG according to the distribution of muscle involvement revealed the following sRAGE concentrations: generalized MG 999.4 ± 90.8 and ocular MG 696.1 ± 161.8 (vs. control; One-way ANOVA: p < 0.001; Post hoc analysis: generalized vs. ocular MG: p = 0.264, generalized MG vs. control: p = 0.008, ocular MG vs. control: p = 0.001). In patients with detectable antibodies specific for acetylcholine receptors (Anti-AChR positive) the sRAGE concentration was 970.0 ± 90.2 compared to those without (seronegative) 670.6 ± 133.1 (vs. control; One-way ANOVA: p < 0.001; Post hoc analysis: Pos vs. Neg.: p = 0.418, Pos vs. control: p = 0.003, Neg. vs. control: p = 0.008). We next investigated the role of RAGE ligands in MG. The concentrations of RAGE ligands in patients with MG and controls were as follows: (HMGB1 [ng/ml] 1.7 ± 0.1 vs. 2.1 ± 0.2; p = 0.058; S100B [pg/ml] 22.5 ± 22.5 vs. 14.4 ± 9.2; p = 0.698; S100A8 [pg/ml] 107.0 ± 59.3 vs. 242.5 ± 103.6; p = 0.347; and AGE-CML [ng/ml] 1100.8 ± 175.1 vs. 1399.8 ± 132.8; p = 0.179).ConclusionsOur data suggest a role for the RAGE pathway in the pathophysiology of MG. Further studies are warranted to elucidate more about this immunological axis in patients with MG.  相似文献   

14.
The 3-keto-l-gulonate 6-phosphate decarboxylase (KGPDC) catalyses the decarboxylation of 3-keto-l-gulonate 6-phosphate to l-xylulose in the presence of magnesium ions. The enzyme is involved in l-ascorbate metabolism and plays an essential role in the pathway of glucuronate interconversion. Crystal structures of Streptococcus mutans KGPDC were determined in the absence and presence of the product analog d-ribulose 5-phosphate. We have observed an 8 Å αB-helix movement and other structural rearrangements around the active site between the apo-structures and product analog bound structure. These drastic conformational changes upon ligand binding are the first observation of this kind for the KGPDC family. The flexibilities of both the α-helix lid and the side chains of Arg144 and Arg197 are associated with substrate binding and product releasing. The open–closed conformational changes of the active site, through the movements of the α-helix lid and the arginine residues are important for substrate binding and catalysis.  相似文献   

15.
Background: There is a growing body of evidence that physical training exerts its potential benefits on the individual health status by modulating the immune system and the whole body metabolism. A better knowledge of the physiologic immune response to exercise may help to understand the benefits of physical exercise in healthy individuals and elite athletes. Aims: This study aims to analyse cardiotrophin-1 (CT-1) and Tumor Necrosis Factor-α (TNF-α) plasma levels at rest and during exercise in elite athletes and healthy controls. Methods: We studied 20 triathletes (TA) and 20 matched controls (CG). Chambers dimensions, left ventricular mass and left ventricular mass index were analysed by echocardiography. VO2 peak and VE/VCO2 were calculated by metabolic stress test. Blood samples were collected before the exercise session, at the exercise peak, and after the end of exercise. ELISA assays were used to measure CT-1 and TNF-α plasma levels. Results:Among TA and CG, no significant differences were found for CT-1 (0.25 ± 0.14 vs 0.20 ± 0.14 fm/l; p = 0.29) and TNF-α (10.8 ± 2.7 vs 9.7 ± 4.0 pm/l; p = 0.29) basal levels. In the TA, plasma levels of CT-1 were significantly different at rest and during exercise (basal 0.25 ± 0.13 pm/l; peak 1.07 ± 1.5 pm/l; post-exercise 0.67 ± 0.77 pm/l; p = 0.04). Conversely, no significant differences were found between basal, peak and post-exercise plasma values of TNF-α (basal 10.8 ± 2.7 pm/l; peak 11.7 ± 2.1 pm/l; post-exercise 11.4 ± 2.5 pm/l; p = 0.78) in TA. Conclusions: This study gives novel insights on the behavior of inflammatory cytokines during physical exercise in athletes and healthy individuals.  相似文献   

16.
We present RIBFIND, a method for detecting flexibility in protein structures via the clustering of secondary structural elements (SSEs) into rigid bodies. To test the usefulness of the method in refining atomic structures within cryoEM density we incorporated it into our flexible fitting protocol (Flex-EM). Our benchmark includes 13 pairs of protein structures in two conformations each, one of which is represented by a corresponding cryoEM map. Refining the structures in simulated and experimental maps at the 5–15 Å resolution range using rigid bodies identified by RIBFIND shows a significant improvement over using individual SSEs as rigid bodies. For the 15 Å resolution simulated maps, using RIBFIND-based rigid bodies improves the initial fits by 40.64% on average, as compared to 26.52% when using individual SSEs. Furthermore, for some test cases we show that at the sub-nanometer resolution range the fits can be further improved by applying a two-stage refinement protocol (using RIBFIND-based refinement followed by an SSE-based refinement). The method is stand-alone and could serve as a general interactive tool for guiding flexible fitting into EM maps.  相似文献   

17.
SHP-1 (PTPN6) is a member of the SHP sub-family of protein tyrosine phosphatases and plays a critical role in the regulation of the JAK/STAT signaling pathway. Previous studies suggested that SHP-1 contains a PTP1B-like second phosphotyrosine pocket that allows for binding of tandem phosphotyrosine residues, such as those found in the activation loop of JAK kinases. To discover the structural nature of the interaction between SHP-1 and the JAK family member, JAK1, we determined the 1.8 Å co-crystal structure of the SHP-1 catalytic domain and a JAK1-derived substrate peptide. This structure reveals electron density for only one bound phosphotyrosine residue. To investigate the role of the predicted second site pocket we determined the structures of SHP-1 in complex with phosphate and sulfate to 1.37 Å and 1.7 Å, respectively, and performed anomalous scattering experiments for a selenate-soaked crystal. These crystallographic data suggest that SHP-1 does not contain a PTP1B-like second site pocket. This conclusion is further supported by analysis of the relative dephosphorylation and binding affinities of mono- and tandem-phosphorylated peptide substrates. The crystal structures instead indicate that SHP-1 contains an extended C-terminal helix α2’ incompatible with the predicted second phosphotyrosine binding site. This study suggests that SHP-1 defines a new category of PTP1B-like protein tyrosine phosphatases with a hindered second phosphotyrosine pocket.  相似文献   

18.
Phosphofructokinase (PFK-1) activity was examined in L3 and adult Teladorsagia circumcincta, both of which exhibit oxygen consumption. Although activities were higher in the adult stage, the kinetic properties of the enzyme were similar in both life cycle stages. T. circumcincta PFK-1 was subject to allosteric inhibition by high ATP concentration, which increased both the Hill coefficient (from 1.4 ± 0.2 to 1.7 ± 0.2 in L3s and 2.0 ± 0.3 to 2.4 ± 0.4 in adults) and the K½ for fructose 6 phosphate (from 0.35 ± 0.02 to 0.75 ± 0.05 mM in L3s and 0.40 ± 0.03 to 0.65 ± 0.05 mM in adults). The inhibitory effects of high ATP concentration could be reversed by fructose 2,6 bisphosphate and AMP, but glucose 1,6 bisphosphate had no effect on activity. Similarly, phosphoenolpyruvate had no effect on activity, while citrate, isocitrate and malate exerted mild inhibitory effects, but only at concentrations exceeding 2 mM. The observed kinetic properties for T. circumcincta PFK-1 were very similar to those reported for purified Ascaris suum PFK-1, though slight differences in sensitivity to ATP concentration suggests there may be subtle variations at the active site. These results are consistent with the conservation of properties of PFK-1 amongst nematode species, despite between species variation in the ability to utilise oxygen.  相似文献   

19.
Warren JJ  Winkler JR  Gray HB 《FEBS letters》2012,586(5):596-602
Redox reactions of tyrosine play key roles in many biological processes, including water oxidation and DNA synthesis. We first review the redox properties of tyrosine (and other phenols) in small molecules and related polypeptides, then report work on (H20)/(Y48)-modified Pseudomonas aeruginosa azurin. The crystal structure of this protein (1.18 Å resolution) shows that H20 is strongly hydrogen bonded to Y48 (2.7–2.8 Å tyrosine-O to histidine-N distance). A firm conclusion is that proper tuning of the tyrosine potential by a proton-accepting base is critical for biological redox functions.  相似文献   

20.
The anti-Alzheimer’s agent galantamine is known to possess anti-amyloid properties. However the exact mechanisms are not clear. We studied the binding interactions of galantamine with amyloid peptide dimer (Aβ1–40) through molecular docking and molecular dynamics simulations. Galantamine’s binding site within the amyloid peptide dimer was identified by docking experiments and the most stable complex was analyzed by molecular dynamics simulation. These studies show that galantamine was interacting with the central region of the amyloid dimer (Lys16–Ala21) and the C-terminal region (Ile31–Val36) with minimum structural drift of Cα atom in those regions. Strikingly, a significant drift was observed at the turn region from Asp23-Gly29 (Cα atom RMSD = 9.2 Å and 11.6 Å at 50 fs and 100 fs respectively). Furthermore, galantamine’s binding mode disrupts the key pi–pi stacking interaction between aromatic rings of Phe19 (chain A) and Phe19 (chain B) and intermolecular hydrogen bonds seen in unbound peptide dimer. Noticeably, the azepine tertiary nitrogen of galantamine was in close proximity to backbone CO of Leu34 (distance <3.5 Å) to stabilize the dimer conformation. In summary, the results indicate that galantamine binding to amyloid peptide dimer leads to a significant conformational change at the turn region (Asp23–Gly29) that disrupts interactions between individual β-strands and promotes a nontoxic conformation of Aβ1–40 to prevent the formation of neurotoxic oligomers.  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