首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到18条相似文献,搜索用时 0 毫秒
1.
The crystal structure of the phospholipase A2 (PLA2) heterodimer from Naja naja sagittifera reveals the presence of a new PLA2-like protein with eight disulphide bridges. The heterodimer is formed between a commonly observed group I PLA2 having seven characteristic disulfide bonds and a novel PLA2-like protein (Cys-PLA2) containing two extra cysteines at two highly conserved sites (positions 32 and 49) of structural and functional importance. The crystals of the heterodimer belong to tetragonal space group P41212 with cell dimensions, a = b = 77.7 A and c = 68.4 A corresponding to a solvent content of 33%, which is one of the lowest values observed so far in the PLA2 crystals. The structure has been solved with molecular replacement method and refined to a final R value of 21.6% [Rfree = 25.6%]. The electron density revealed the presence of cysteines 32 and 49 that are covalently linked to give rise to an eighth disulphide bridge in the PLA2-like monomer. A non-protein high-quality electron density was also observed at the substrate-binding site in the PLA2-like protein that has been interpreted as N-acetylglucosamine. The overall tertiary folds of the two monomers are similar having all features of PLA2-type folding. A zinc ion is detected at the interface of the heterodimer with fivefold coordination while another zinc ion was found on the surface of Cys-PLA2 with sixfold coordination. The conformations of the calcium-binding loops of both monomers are significantly different from each other as well as from those in other group I PLA2s. The N-acetylglucosamine molecule is favorably placed in the substrate-binding site of Cys-PLA2 and forms five hydrogen bonds and several van der Waals interactions with protein atoms, thus indicating a strong affinity. It also provides clue of the possible mechanism of sugar recognition by PLA2 and PLA2-like proteins. The formation of heterodimer seems to have been induced by zinc ion.  相似文献   

2.
Lactoperoxidase (LPO) belongs to mammalian heme peroxidase superfamily, which also includes myeloperoxidase (MPO), eosinophil peroxidase (EPO), and thyroid peroxidase (TPO). LPO catalyzes the oxidation of a number of substrates including thiocyanate while TPO catalyzes the biosynthesis of thyroid hormones. LPO is also been shown to catalyze the biosynthesis of thyroid hormones indicating similar functional and structural properties. The binding studies showed that 2‐mercaptoimidazole (MZY) bound to LPO with a dissociation constant of 0.63 µM. The inhibition studies showed that the value of IC50 was 17 µM. The crystal structure of the complex of LPO with MZY showed that MZY bound to LPO in the substrate‐binding site on the distal heme side. MZY was oriented in the substrate‐binding site in such a way that the sulfur atom is at a distance of 2.58 Å from the heme iron. Previously, a similar compound, 3‐amino‐1,2,4‐triazole (amitrole) was also shown to bind to LPO in the substrate‐binding site on the distal heme side. The amino nitrogen atom of amitrole occupied the same position as that of sulfur atom in the present structure indicating a similar mode of binding. Recently, the structure of the complex of LPO with a potent antithyroid drug, 1‐methylimidazole‐2‐thiol (methimazole, MMZ) was also determined. It showed that MMZ bound to LPO in the substrate‐binding site on the distal heme side with 2 orientations. The position of methyl group was same in the 2 orientations while the positions of sulfur atom differed indicating a higher preference for a methyl group.  相似文献   

3.
The acyl‐AMP forming family of adenylating enzymes catalyze two‐step reactions to activate a carboxylate with the chemical energy derived from ATP hydrolysis. X‐ray crystal structures have been determined for multiple members of this family and, together with biochemical studies, provide insights into the active site and catalytic mechanisms used by these enzymes. These studies have shown that the enzymes use a domain rotation of 140° to reconfigure a single active site to catalyze the two partial reactions. We present here the crystal structure of a new medium chain acyl‐CoA synthetase from Methanosarcina acetivorans. The binding pocket for the three substrates is analyzed, with many conserved residues present in the AMP binding pocket. The CoA binding pocket is compared to the pockets of both acetyl‐CoA synthetase and 4‐chlorobenzoate:CoA ligase. Most interestingly, the acyl‐binding pocket of the new structure is compared with other acyl‐ and aryl‐CoA synthetases. A comparison of the acyl‐binding pocket of the acyl‐CoA synthetase from M. acetivorans with other structures identifies a shallow pocket that is used to bind the medium chain carboxylates. These insights emphasize the high sequence and structural diversity among this family in the area of the acyl‐binding pocket. Proteins 2009. © 2009 Wiley‐Liss, Inc.  相似文献   

4.
The synthetic peptide Z‐(Aib)10‐OH was crystallized from hot methanol by slow evaporation. The crystal used for data collection reflected synchrotron radiation to sub‐atomic resolution, where the bonding electron density becomes visible between the non‐hydrogen atoms. Crystals belong to the centrosymmetric space group P . Both molecules in the asymmetric unit form regular 310‐helices. All residues in each molecule possess the same handedness, which is in contrast to all other crystal structure determined to date of longer Aib‐homopeptides. These other peptides are C‐terminal protected by OtBu or OMe. In these cases, because of the missing ability of the C‐terminal protection group to form a hydrogen bond to the residue i‐3, the sense of the helix is reversed in the last residue. Here, the C‐terminal OH‐groups form hydrogen bonds to the residues i‐3, in part mediated by water molecules. This makes Z‐(Aib)10‐OH an Aib‐homopeptide with three complete 310‐helical turns in spite of the shorter length it has compared with Z‐(Aib)11‐OtBu, the only homopeptide to date with three complete turns.  相似文献   

5.
6.
The phosphoserine/threonine binding protein 14‐3‐3 stimulates the catalytic activity of protein kinase C‐ε (PKCε) by engaging two tandem phosphoserine‐containing motifs located between the PKCε regulatory and catalytic domains (V3 region). Interaction between 14‐3‐3 and this region of PKCε is essential for the completion of cytokinesis. Here, we report the crystal structure of 14‐3‐3ζ bound to a synthetic diphosphorylated PKCε V3 region revealing how a consensus 14‐3‐3 site and a divergent 14‐3‐3 site cooperate to bind to 14‐3‐3 and so activate PKCε. Thermodynamic data show a markedly enhanced binding affinity for two‐site phosphopeptides over single‐site 14‐3‐3 binding motifs and identifies Ser 368 as a gatekeeper phosphorylation site in this physiologically relevant 14‐3‐3 ligand. This dual‐site intra‐chain recognition has implications for other 14‐3‐3 targets, which seem to have only a single 14‐3‐3 motif, as other lower affinity and cryptic 14‐3‐3 gatekeeper sites might exist.  相似文献   

7.
Thymidine‐3′‐monophosphate (3′‐TMP) is a competitive inhibitor analogue of the 3′‐CMP and 3′‐UMP natural product inhibitors of bovine pancreatic ribonuclease A (RNase A). Isothermal titration calorimetry experiments show that 3′‐TMP binds the enzyme with a dissociation constant (Kd) of 15 μM making it one of the strongest binding members of the five natural bases found in nucleic acids (A, C, G, T, and U). To further investigate the molecular properties of this potent natural affinity, we have determined the crystal structure of bovine pancreatic RNase A in complex with 3′‐TMP at 1.55 Å resolution and we have performed NMR binding experiments with 3′‐CMP and 3′‐TMP. Our results show that binding of 3′‐TMP is very similar to other natural and non‐natural pyrimidine ligands, demonstrating that single nucleotide affinity is independent of the presence or absence of a 2′‐hydroxyl on the ribose moiety of pyrimidines and suggesting that the pyrimidine binding subsite of RNase A is not a significant contributor of inhibitor discrimination. Accumulating evidence suggests that very subtle structural, chemical, and potentially motional variations contribute to ligand discrimination in this enzyme. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

8.
The anti‐lipopolysaccharide factor ALF‐Pm3 is a 98‐residue protein identified in hemocytes from the black tiger shrimp Penaeus monodon. It was expressed in Pichia pastoris from the constitutive glyceraldehyde‐3‐phosphate dehydrogenase promoter as a folded and 15N uniformly labeled rALF‐Pm3 protein. Its 3D structure was established by NMR and consists of three α‐helices packed against a four‐stranded β‐sheet. The C34? C55 disulfide bond was shown to be essential for the structure stability. By using surface plasmon resonance, we demonstrated that rALF‐Pm3 binds to LPS, lipid A and to OM®‐174, a soluble analogue of lipid A. Biophysical studies of rALF‐Pm3/LPS and rALF‐Pm3/OM®‐174 complexes indicated rather high molecular sized aggregates, which prevented us to experimentally determine by NMR the binding mode of these lipids to rALF‐Pm3. However, on the basis of striking structural similarities to the FhuA/LPS complex, we designed an original model of the possible lipid A‐binding site of ALF‐Pm3. Such a binding site, located on the ALF‐Pm3 β‐sheet and involving seven charged residues, is well conserved in ALF‐L from Limulus polyphemus and in ALF‐T from Tachypleus tridentatus. In addition, our model is in agreement with experiments showing that β‐hairpin synthetic peptides corresponding to ALF‐L β‐sheet bind to LPS. Delineating lipid A‐binding site of ALFs will help go further in the de novo design of new antibacterial or LPS‐neutralizing drugs. © 2008 Wiley Periodicals, Inc. Biopolymers 91: 207–220, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

9.
ß‐Amyloid (Aß) immunotherapy has become a promising strategy for reducing the level of Aß in brain. New immunological approaches have been recently proposed for rapid, early diagnosis, and molecular treatment of neurodegenerative diseases related to Alzheimer's Disease (AD). The combination of proteolytic epitope excision and extraction and mass spectrometry using digestion with various proteases has been shown to be an efficient tool for the identification and molecular characterization of antigenic determinants. Here, we report the identification of the Aβ epitope recognized by the variable domain of single chain llama anti‐Aβ‐antibodies, termed Aβ‐nanobodies, that have been discovered in the blood of camelids and found to be promising candidates for immunotherapy of AD. The epitope recognized by two Aβ‐specific nanobodies was identified by proteolytic epitope extraction‐ and excision‐mass spectrometry using a series of proteases (trypsin, chymotrypsin, GluC‐protease, and LysC‐protease). Matrix‐assisted laser desorption ionization – mass spectrometric analysis of the affinity – elution fraction provided the epitope, Aβ(17–28), in the mid‐ to carboxy‐terminal domain of Aβ, which has been shown to exert an Aß‐fibril inhibiting effect. Affinity studies of the synthetic epitope confirmed that the Aβ(17–28) peptide is the minimal fragment that binds to the nanobodies. The interactions between the nanobodies and full length Aβ(1–40) or Aβ‐peptides containing or lacking the epitope sequence were further characterized by enzyme linked immunosorbent assay and bioaffinity analysis. Determinations of binding affinities between the Aβ‐nanobodies and Aβ(1–40) and the Aβ(17–28) epitope provided KD values of approximately 150 and 700 nmol, respectively. Thus, the knowledge of the epitope may be highly useful for future studies of Aβ‐aggregation (oligomerization and fibril formation) and for designing new aggregation inhibitors. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

10.
Two complexes of Tb3+, Gd3+/Tb3+ and one heteronuclear crystal Gd3+/Tb3+ with phenoxyacetic acid (HPOA) and 2,4,6‐tris‐(2‐pyridyl)‐s–triazine (TPTZ) have been synthesized. Elemental analysis, rare earth coordination titration, inductively coupled plasma atomic emission spectrometry (ICP‐AES) and thermogravimetric analysis‐differential scanning calorimetry (TG‐DSC) analysis show that the two complexes are Tb2(POA)6(TPTZ)2·6H2O and TbGd(POA)6(TPTZ)2·6H2O, respectively. The crystal structure of TbGd(POA)6(TPTZ)2·2CH3OH was determined using single‐crystal X‐ray diffraction. The monocrystal belongs to the triclinic system with the P‐1 space group. In particular, each metal ion is coordinately bonded to three nitrogen atoms of one TPTZ and seven oxygen atoms of three phenoxyacetic ions. Furthermore, there exist two coordinate forms between C6H5OCH2COO and the metal ions in the crystal. One is a chelating bidentate, the other is chelating and bridge coordinating. Fluorescence determination shows that the two complexes possess strong fluorescence emissions. Furthermore, the fluorescence intensity of the Gd3+/Tb3+ complex is much stronger than that of the undoped complex, which may result from a decrease in the concentration quench of Tb3+ ions, and intramolecular energy transfer from the ligands coordinated with Gd3+ ions to Tb3+ ions. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

11.
The heart‐specific isoform of 6‐phosphofructo‐2‐kinase/fructose‐2,6‐bisphosphatase (PFKFB2) is an important regulator of glycolytic flux in cardiac cells. Here, we present the crystal structures of two PFKFB2 orthologues, human and bovine, at resolutions of 2.0 and 1.8 Å, respectively. Citrate, a TCA cycle intermediate and well‐known inhibitor of PFKFB2, co‐crystallized in the 2‐kinase domains of both orthologues, occupying the fructose‐6‐phosphate binding‐site and extending into the γ‐phosphate binding pocket of ATP. This steric and electrostatic occlusion of the γ‐phosphate site by citrate proved highly consequential to the binding of co‐complexed ATP analogues. The bovine structure, which co‐crystallized with ADP, closely resembled the overall structure of other PFKFB isoforms, with ADP mimicking the catalytic binding mode of ATP. The human structure, on the other hand, co‐complexed with AMPPNP, which, unlike ADP, contains a γ‐phosphate. The presence of this γ‐phosphate made adoption of the catalytic ATP binding mode impossible for AMPPNP, forcing the analogue to bind atypically with concomitant conformational changes to the ATP binding‐pocket. Inhibition kinetics were used to validate the structural observations, confirming citrate's inhibition mechanism as competitive for F6P and noncompetitive for ATP. Together, these structural and kinetic data establish a molecular basis for citrate's negative feed‐back loop of the glycolytic pathway via PFKFB2. Proteins 2016; 85:117–124. © 2016 Wiley Periodicals, Inc.  相似文献   

12.
The single‐crystal structure of the collagen‐like peptide (Pro‐Pro‐Gly)4‐Hyp‐Asp‐Gly‐(Pro‐Pro‐Gly)4, was analyzed at 1.02 Å resolution. The overall average helical twist (θ = 49.6°) suggests that this peptide adopts a 7/2 triple‐helical structure and that its conformation is very similar to that of (Gly‐Pro‐Hyp)9, which has the typical repeating sequence in collagen. High‐resolution studies on other collagen‐like peptides have shown that imino acid‐rich sequences preferentially adopt a 7/2 triple‐helical structure (θ = 51.4°), whereas imino acid‐lean sequences adopt relaxed conformations (θ < 51.4°). The guest Gly‐Hyp‐Asp sequence in the present peptide, however, has a large helical twist (θ = 61.1°), whereas that of the host Pro‐Pro‐Gly sequence is small (θ = 46.7°), indicating that the relationship between the helical conformation and the amino acid sequence of such peptides is complex. In the present structure, a strong intermolecular hydrogen bond between two Asp residues on the A and B strands might induce the large helical twist of the guest sequence; this is compensated by a reduced helical twist in the host, so that an overall 7/2‐helical symmetry is maintained. The Asp residue in the C strand might interact electrostatically with the N‐terminus of an adjacent molecule, causing axial displacement, reminiscent of the D‐staggered structure in fibrous collagens. © 2013 Wiley Periodicals, Inc. Biopolymers 99: 436–447, 2013.  相似文献   

13.
The force driving the conversion from the acyl intermediate to the tetrahedral intermediate in the deacylation reaction of serine proteases remains unclear. The crystal structure of 6‐guanidinohexanoyl trypsin was determined at pH 7.0, near the optimum reaction pH, at 1.94 Å resolution. In this structure, three water molecules are observed around the catalytic site. One acts as a nucleophile to attack the acyl carbonyl carbon while the other two waters fix the position of the catalytic water through a hydrogen bond. When the acyl carbonyl oxygen oscillates thermally, the water assumes an appropriate angle to catalyze the deacylation. Proteins 2013. © 2012 Wiley Periodicals, Inc.  相似文献   

14.
[Ni(C11H9N2O5)2(H2O)2]?3(C3H7NO) ( 1 ) and [Co(C11H9N2O5)2(H2O)2]?3(C3H7NO) ( 2 ) are synthesized and characterized by elemental analysis, FT‐IR spectra, magnetic susceptibility, and thermal analysis. In addition, the crystal structure of Ni(II) complex is presented. Both complexes show distorted octahedral geometry. In 1 and 2, metal ions are coordinated by two oxygen atoms of salicylic residue and two nitrogen atoms of maleic amide residue from two ligands, and two oxygen atoms from two water molecules. In this paper, both compounds showed excellent inhibitory effects against human carbonic anhydrase (hCA) isoforms I, and II, α‐glycosidase, acetylcholinesterase (AChE), and butyrylcholinesterase (BChE). Compounds 1 and 2 had Ki values of 18.36 ± 4.38 and 26.61 ± 7.54 nM against hCA I and 13.81 ± 3.02 and 29.56 ± 6.52 nM against hCA II, respectively. On the other hand, their Ki values were found to be 487.45 ± 54.18 and 453.81 ± 118.61 nM against AChE and 199.21 ± 50.35 and 409.41 ± 6.86 nM against BChE, respectively.  相似文献   

15.
This is the first structural evidence of alpha-tocopherol (alpha-TP) as a possible candidate against inflammation, as it inhibits phospholipase A2 specifically and effectively. The crystal structure of the complex formed between Vipera russelli phospholipase A2 and alpha-tocopherol has been determined and refined to a resolution of 1.8 A. The structure contains two molecules, A and B, of phospholipase A2 in the asymmetric unit, together with one alpha-tocopherol molecule, which is bound specifically to one of them. The phospholipase A2 molecules interact extensively with each other in the crystalline state. The two molecules were found in a stable association in the solution state as well, thus indicating their inherent tendency to remain together as a structural unit, leading to significant functional implications. In the crystal structure, the most important difference between the conformations of two molecules as a result of their association pertains to the orientation of Trp31. It may be noted that Trp31 is located at the mouth of the hydrophobic channel that forms the binding domain of the enzyme. The values of torsion angles (phi, psi, chi(1) and chi(2)) for both the backbone as well as for the side-chain of Trp31 in molecules A and B are -94 degrees, -30 degrees, -66 degrees, 116 degrees and -128 degrees, 170 degrees, -63 degrees, -81 degrees, respectively. The conformation of Trp31 in molecule A is suitable for binding, while that in B hinders the passage of the ligand to the binding site. Consequently, alpha-tocopherol is able to bind to molecule A only, while the binding site of molecule B contains three water molecules. In the complex, the aromatic moiety of alpha-tocopherol is placed in the large space at the active site of the enzyme, while the long hydrophobic channel in the enzyme is filled by hydrocarbon chain of alpha-tocopherol. The critical interactions between the enzyme and alpha-tocopherol are generated between the hydroxyl group of the six-membered ring of alpha-tocopherol and His48 N(delta1) and Asp49 O(delta1) as characteristic hydrogen bonds. The remaining part of alpha-tocopherol interacts extensively with the residues of the hydrophobic channel of the enzyme, giving rise to a number of hydrophobic interactions, resulting in the formation of a stable complex.  相似文献   

16.
To clarify the higher eukaryotic initiation factor 4E (eIF4E) binding selectivity of 4E‐binding protein 2 (4E‐BP2) than of 4E‐BP1, as determined by Trp fluorescence analysis, the crystal structure of the eIF4E binding region of 4E‐BP2 in complex with m7GTP‐bound human eIF4E has been determined by X‐ray diffraction analysis and compared with that of 4E‐BP1. The crystal structure revealed that the Pro47‐Ser65 moiety of 4E‐BP2 adopts a L ‐shaped conformation involving extended and α‐helical structures and extends over the N‐terminal loop and two different helix regions of eIF4E through hydrogen bonds, and electrostatic and hydrophobic interactions; these features were similarly observed for 4E‐BP1. Although the pattern of the overall interaction of 4E‐BP2 with eIF4E was similar to that of 4E‐BP1, a notable difference was observed for the 60–63 sequence in relation to the conformation and binding selectivity of the 4E‐BP isoform, i.e. Met‐Glu‐Cys‐Arg for 4E‐BP1 and Leu‐Asp‐Arg‐Arg for 4E‐BP2. In this paper, we report that the structural scaffold of the eIF4E binding preference for 4E‐BP2 over 4E‐BP1 is based on the stacking of the Arg63 planar side chain on the Trp73 indole ring of eIF4E and the construction of a compact hydrophobic space around the Trp73 indole ring by the Leu59‐Leu60 sequence of 4E‐BP2. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

17.
18.
Transient receptor potential melastatin 2 (TRPM2) is a Ca2+‐permeable, nonselective cation channel involved in diverse physiological processes such as immune response, apoptosis, and body temperature sensing. TRPM2 is activated by ADP‐ribose (ADPR) and 2′‐deoxy‐ADPR in a Ca2+‐dependent manner. While two distinct binding sites exist for ADPR that exert different functions dependent on the species, the involvement of either binding site regarding the superagonistic effect of 2′‐deoxy‐ADPR is not clear yet. Here, we report the crystal structure of the MHR1/2 domain of TRPM2 from zebrafish (Danio rerio), and show that both ligands bind to this domain and activate the channel. We identified a so far unrecognized Zn2+‐binding domain that was not resolved in previous cryo‐EM structures and that is conserved in most TRPM channels. In combination with patch clamp experiments we comprehensively characterize the effect of the Zn2+‐binding domain on TRPM2 activation. Our results provide insight into a conserved motif essential for structural integrity and channel activity.  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