首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Nonenzymatic glycation of antithrombin III has been reported to cause the reduction of heparin-catalyzed thrombin-inhibiting activity in diabetes. The effect of in vitro nonenzymatic glycation of pure antithrombin III on heparin binding and heparin-potentiated activity under a variety of buffers and pH values was studied to further clarify the physiological significance of this reaction. The extent of glycation, measured by the fructosamine assay and [14C]glucose binding, was enhanced by the presence of phosphate ion (pH 7.45, 8.5 and 9.5) and increased linearly with increasing phosphate ion concentration from 0.01 to 0.2 M phosphate. Conversely, the heparin-catalyzed antithrombin activity decreased from 93.1% of controls for 0.01 M phosphate to 73.5% for 0.2 M phosphate as the extent of glycation increased. The increase in intrinsic fluorescence induced by binding of heparin to antithrombin III was also moderated by glycation of antithrombin III in a dose-dependent manner with a negative correlation coefficient of -0.94. Direct measurement of the heparin binding by affinity chromatography showed a decrease in the heparin-binding fraction which correlated with the degree of glycation and the decrease in heparin-catalyzed activity. These studies suggest that nonenzymatic glycation may be responsible for the reduction in antithrombin III activity observed in some diabetics.  相似文献   

2.
The present study deals with the conformation in solution of two heparin octasaccharides containing the pentasaccharide sequence GlcN(NAc,6S)-GlcA-GlcN(NS,3,6S)-IdoA(2S)-GlcN(NS,6S) [AGA*IA; where GlcN(NAc,6S) is N-acetylated, 6-O-sulfated alpha-D-glucosamine, GlcN(NS,3,6S) is N,3,6-O-trisulfated alpha-D-glucosamine and IdoA(2S) is 2-O-sulfated IdoA (alpha-L-iduronic acid)] located at different positions in the heparin chain and focuses on establishing geometries of IdoA residues (IdoA(2S) and IdoA) both inside and outside the AGA*IA sequence. AGA*IA constitutes the active site for AT (antithrombin) and is essential for the expression of high anticoagulant and antithrombotic activities. Analysis of NMR parameters [NOEs (nuclear Overhauser effects), transferred NOEs and coupling constants] for the two octasaccharides indicated that between the 1C4 and 2S0 conformations present in dynamic equilibrium in the free state for the IdoA(2S) residue within AGA*IA, AT selects the 2S0 form, as previously shown [Hricovini, Guerrini, Bisio, Torri, Petitou and Casu (2001) Biochem. J. 359, 265-272]. Notably, the 2S0 conformation is also adopted by the non-sulfated IdoA residue preceding AGA*IA that, in the absence of AT, adopts predominantly the 1C4 form. These results further support the concept that heparin-binding proteins influence the conformational equilibrium of iduronic acid residues that are directly or indirectly involved in binding and select one of their equi-energetic conformations for best fitting in the complex. The complete reversal of an iduronic acid conformation preferred in the free state is also demonstrated for the first time. Preliminary docking studies provided information on the octasaccharide binding location agreeing most closely with the experimental data. These results suggest a possible biological role for the non-sulfated IdoA residue preceding AGA*IA, previously thought not to influence the AT-binding properties of the pentasaccharide. Thus, for each AT binding sequence longer than AGA*IA, the interactions with the protein could differ and give to each heparin fragment a specific biological response.  相似文献   

3.
Cesaretti M  Luppi E  Maccari F  Volpi N 《Glycobiology》2004,14(12):1275-1284
Heparin with high anticoagulant activity (activated partial thromboplastin time of 347 +/- 56.4 and anti-Xa activity of 317 +/- 48.3) was isolated from the marine clam species Tapes phylippinarum in an amount of approximately 2.1 mg/g dry animals. Agarose-gel electrophoresis showed a high content of the slow-moving heparin component (22 +/- 6.8%) and 78 +/- 5.4% of the fast-moving species. An average molecular mass of 13,600 was calculated by PAGE analysis, whereas a number average molecular weight Mn value of 10,700, a weight average molecular weight Mw of 14,900, and a dispersity index Mn/Mw of 1.386 were obtained by high-performance size-exclusion chromatography. Structural analysis of clam heparin, performed by depolymerizing heparin samples with heparinase (EC 4.2.2.7) and then separating the resulting unsaturated oligosaccharides by strong anion exchange-HPLC revealed the presence of large amounts (more than 130% than standard pharmaceutical heparin obtained from bovine intestine) of the oligosaccharide sequence bearing part of the ATIII-binding region, DeltaUA2S (1-->4)-alpha-D-GlcN2S6S (1-->4)-alpha-L-IdoA (1-->4)-alpha-D-GlcNAc6S (1-->4)-beta-D-GlcA (1-->4)-alpha-D-GlcN2S3S6S in the T. phylippinarum heparin, in comparison with bovine mucosal heparin and a sample of porcine mucosal heparin previously published. Furthermore, as expected from the oligosaccharide compositional analysis, due to the presence of a great mol % (80.6%) of the trisulfated disaccharide DeltaUA2S(1-->4)-alpha-D-GlcN2S6S, mollusc heparin is a more sulfated polysaccharide than bovine mucosal heparin (73.5%) and a sample of porcine mucosal (72.8%) heparin previously reported. To our knowledge, this is the first article describing a clam heparin having the ATIII binding site mainly identical to that of human and porcine intestinal mucosal heparins and bovine intestinal mucosal heparin but different from that found in beef lung heparin.  相似文献   

4.
The importance of 3-O- and 6-O-sulfated glucosamine residues within the heparin octasaccharide iduronic acid(1)----N-acetylglucosamine 6-O-sulfate(2)----glucuronic acid(3)----N-sulfated glucosamine 3,6-di-O-sulfate(4)----iduronic acid 2-O-sulfate(5)----N-sulfated glucosamine 6-O-sulfate(6)----iduronic acid 2-O-sulfate(7)----anhydromannitol 6-O-sulfate(8) was determined by comparing with synthetic tetra- and penta-saccharides its ability to bind human antithrombin. The octasaccharide had an affinity for antithrombin of 1 X 10(-8) M (10.2 kcal/mol) measured by intrinsic fluorescence enhancement at 6 degrees C. The synthetic pentasaccharide, consisting of residues 2-6, had an affinity of 3 X 10(-8) M (9.6 kcal/mol). The same pentasaccharide, except lacking the 3-O-sulfate on residue 4, had an affinity of 5 X 10(-4) M (4.5 kcal/mol) measured by equilibrium dialysis. The tetrasaccharide, consisting of residues 2-5, bound antithrombin with an affinity of 5 X 10(-6) M (6.8 kcal/mol). The tetrasaccharide, consisting of residues 3-6, had an affinity of 5 X 10(-5) M (5.5 kcal/mol). Since the loss of either the 6-O-sulfated residue 2 or the 3-O-sulfate of residue 4 results in a 4-5 kcal/mol or a 40-50% loss in binding energy of the pentasaccharide, these two residues must be the major contributors to the binding and must be linked to the biologic activity of the octasaccharide.  相似文献   

5.
Chemical modifications have demonstrated that the ultraviolet difference spectrum produced when heparin interacts with antithrombin III is due primarily to changes in the tryptophan environment. This is based on the observation that this spectrum could be abolished by treatment of antithrombin III with dimethyl (2-hydroxy-5-nitrobenzyl) sulfonium bromide but not with tetranitromethane. The tryptophan-modified antithrombin III is still capable of binding to thrombin even when it has lost 85% of heparin cofactor activity. A marked decrease in reactivity of tryptophan residues is observed when modification is carried out in the presence of heparin. Evidence is presented that tryptophan is in the heparin binding site.  相似文献   

6.
The presence of two unfolding domains in antithrombin III during its denaturation in guanidinium chloride has previously been reported (Villanueva, G. B., and Allen, N. (1983) J. Biol. Chem. 258, 11010-11013). In the present work, we report the results of refolding studies on antithrombin III. Circular dichroism and intrinsic fluorescence studies have demonstrated that the first unfolding domain of low stability (midpoint at 0.7 M guanidinium chloride) is irreversible upon renaturation, whereas the second unfolding domain (midpoint at 2.3 M guanidinium chloride) is reversible. The intermediate form of antithrombin III, termed AT-IIIR, which has lost the structural features of the first domain was investigated. Clotting assays and electrophoretic analyses showed that AT-IIIR had lost 60% of heparin cofactor activity but was still capable of forming sodium dodecyl sulfate-stable complexes with thrombin. Although certain regions of this molecule do not refold to the conformation of native antithrombin III, the tryptophan residues refold to a conformation identical with the native state. This was demonstrated by fluorescence quenching, solvent perturbation, and chemical modification studies. However, the tryptophan-ascribed fluorescence enhancement and absorption difference spectrum which occur when heparin binds to antithrombin III are reduced by 70%. On the basis of these data, the binding of heparin to antithrombin III is interpreted in terms of a two-step mechanism. The primary binding occurs in the region without tryptophan and is followed by a secondary conformational rearrangement which affects the tryptophan environment. The mechanism of the binding of heparin and antithrombin III has been previously studied by kinetic methods, and the data also support a two-step mechanism. The agreement of these two studies employing entirely different approaches to the same problem lends support to the validity of this postulated mechanism.  相似文献   

7.
Heparin and low-molecular-weight heparins (LMWHs) are anticoagulant drugs that mainly inhibit the coagulation cascade by indirectly interacting with factor Xa and factor IIa (thrombin). Inhibition of factor Xa by antithrombin (AT) requires the activation of AT by specific pentasaccharide sequences containing 3-O-sulfated glucosamine. Activated AT also inhibits thrombin by forming a stable ternary complex of AT, thrombin, and a polysaccharide (requires at least an 18-mer/octadeca-mer polysaccharide). The full structure of any naturally occurring octadecasaccharide sequence has yet to be determined. In the context of the development of LMWH biosimilars, structural data on such important biological mediators could be helpful for better understanding and regulatory handling of these drugs. Here we present the isolation and identification of an octadecasaccharide with very high anti-factor Xa activity (∼3 times higher than USP [U.S. Pharmacopeia] heparin). The octadecasaccharide was purified using five sequential chromatographic methods with orthogonal specificity, including gel permeation, AT affinity, strong anion exchange, and ion-pair chromatography. The structure of the octadecasaccharide was determined by controlled enzymatic sequencing and nuclear magnetic resonance (NMR). The isolated octadecasaccharide contained three consecutive AT-binding sites and was tested in coagulation assays to determine its biological activity. The isolation of this octadecasaccharide provides new insights into the modulation of thrombin activity.  相似文献   

8.
Tri-calciumphosphate was found to have not only a known adsorption capacity for factors of the prothrombin complex, but also for antithrombin III. Depending on the inserted blood stabilizer the human plasma fractions Cohn I, PPSB and antithrombin III may be isolated from the same initial material in the area of the transfusion service. Enriching antithrombin III is achieved by a three-stage procedure under aseptic conditions in a closed system. Liberating antithrombin III from calciumphosphate is made with care without using any concentrated salt solutions.  相似文献   

9.
A peptide model for the heparin binding site of antithrombin III (ATIII) was synthesized to elucidate the structural consequences of heparin binding. This peptide [ATIII(123-139)] and a sequence-permuted analogue (ATIII random) showed similar conformational behavior (as analyzed by circular dichroism spectroscopy) in aqueous and organic media. In the presence of heparin, however, the peptide ATIII(123-139) assumed a stable conformation, whereas peptide ATIII random did not. Complex formation was saturable and sensitive to salt. The ATIII(123-139)-heparin complex contained beta-structure, rather than helical structure. This finding is incompatible with current models of heparin binding and suggests that heparin binding may induce nonnative structures at the binding site which could, in turn, lead to activation of ATIII. The peptide ATIII(123-139) was able to inhibit the binding of ATIII by heparin, consistent with the notion that this peptide may be a model for the heparin binding site.  相似文献   

10.
Naimy H  Leymarie N  Bowman MJ  Zaia J 《Biochemistry》2008,47(10):3155-3161
Heparan sulfate (HS) is a sulfated glycosaminoglycan attached to a core protein on the cell surface. Protein binding to cell surface HS is a key regulatory event for many cellular processes such as blood coagulation, cell proliferation, and migration. The concept whereby protein binding to HS is not random but requires a limited number of sulfation patterns is becoming clear. Here we describe a hydrophobic trapping assay for screening a library of heparin hexasaccharides for binders to antithrombin III (ATIII). The hexasaccharide compositions are defined with their building block content in the following format: (DeltaHexA:HexA:GlcN:SO 3:Ac). Of five initial compositions present in the library, (1:2:3:6:1), (1:2:3:7:1), (1:2:3:7:0), (1:2:3:8:0), and (1:2:3:9:0), only two are shown to bind ATIII, namely, (1:2:3:8:0) and (1:2:3:9:0). The use of amide hydrophilic interaction (HILIC) liquid chromatography-mass spectrometry permitted reproducible quantitative analysis of the composition of the initial library as well as that of the binding fraction. The specificity of the hexasaccharides binding ATIII was confirmed by assaying their ability to enhance ATIII-mediated inhibition of Factor Xa in vitro.  相似文献   

11.
12.
The molecular basis and functional properties of a variant antithrombin (AT) protein. AT Budapest 3, were studied. A single base substitution was identified in codon 99, TC→ TC, altering the normal leucine to phenylalanine. The proband presented with a history of venous thrombotic disease and was found to be homozygous for the mutation. The variant protein demonstrated reduced heparin affinity and reduced antiproteinase activity in the presence of either unfractionated heparin or the AT-binding heparin pentasaccharide, when compared to normal AT. A small change in the isoelectric point was also identified. The substituted amino acid residue of AT Budapest 3 is located near to the proposed AT heparin binding site, and it is suggested that reduced heparin affinity of the variant protein may result from substitution-induced distortion of positive charge geometry in the binding site and/or changes in its position relative to the rest of the inhibitor molecule.  相似文献   

13.
Antithrombin (AT), a serine protease inhibitor, circulates in blood in two major isoforms, α and β, which differ in their amount of glycosylation and affinity for heparin. After binding to this glycosaminoglycan, the native AT conformation, relatively inactive as a protease inhibitor, is converted to an activated form. In this process, β‐AT presents the higher affinity for heparin, being suggested as the major AT glycoform inhibitor in vivo. However, either the molecular basis demonstrating the differences in heparin binding to both AT isoforms or the mechanism of its conformational activation are not fully understood. Thus, the present work evaluated the effects of glycosylation and heparin binding on AT structure, function, and dynamics. Based on the obtained data, besides the native and activated forms of AT, an intermediate state, previously proposed to exist between such conformations, was also spontaneously observed in solution. Additionally, Asn135‐linked oligosaccharide caused a bending in AT‐bounded heparin, moving such polysaccharide away from helix D, which supports its reduced affinity for α‐AT. The obtained data supported the proposal of an atomic‐level, solvent and amino acid residues accounting, putative model for the transmission of the conformational signal from heparin binding exosite to β‐sheet A and the reactive center loop, also supporting the identification of differences in such transmission between the serpin glycoforms involving helix D, where the Asn135‐linked oligosaccharide stands. Such intramolecular rearrangements, together with heparin dynamics over AT surface, may support an atomic‐level explanation for the Asn135‐linked glycan influence over heparin binding and AT activation. Proteins 2011; © 2011 Wiley‐Liss, Inc.  相似文献   

14.
The conformation in solution of the pentasaccharide methyl glycoside (As-G-A*-Is-AM; 1), which represents the binding site of heparin for Antithrombin III, has been investigated using molecular mechanics and 1H-n.m.r. spectroscopy. The pentasaccharide has a rather rigid (As-G-A*) and a more flexible (Is-AM) region. A simplified model of 1, comprising two conformations, corresponding to the 1C4 and the 2S0 forms of the iduronate residue, and modified at the G-A* glycosidic linkage with respect to the energy minimum, reproduces most of the observed 3J values and n.O.e. enhancements. The possible role in the binding to Antithrombin III of a low-energy conformer, not observed in solution, is discussed.  相似文献   

15.
16.
Highly active heparin species with multiple binding sites for antithrombin   总被引:4,自引:0,他引:4  
Porcine heparin has been fractionated by Sephadex G-100 gel filtration and affinity chromatography into mucopolysaccharide species with approximate molecular sizes of 20,000 daltons, and 7000 daltons, respectively. The larger component has a specific anticoagulant activity of 738 USP units/mg and contains two binding regions for antithrombin. The smaller component has a specific anticoagulant activity of 363 USP units/mg and possesses only a single interaction site for the inhibitor. These results provide the first demonstration that heparin molecules may bear multiple binding sites for antithrombin.  相似文献   

17.
We have isolated from nitrous acid cleavage products of heparin two major octasaccharide fragments which bind with high affinity to human antithrombin. Octasaccharide S, with the predominant structure iduronic acid----N-acetylglucosamine 6-O-sulfate----glucuronic acid-----N-sulfated glucosamine 3,6-di-O-sulfate----iduronic acid 2-O-sulfate----N-sulfated glucosamine 6-O-sulfate----iduronic acid 2-O-sulfate----anhydromannitol 6-O-sulfate, is sensitive to cleavage by Flavobacterium heparinase as well as platelet heparitinase and binds to antithrombin with a dissociation constant of (5-15) X 10(-8) M. Octasaccharide R, with the predominant structure iduronic acid 2-O-sulfate----N-sulfated glucosamine 6-O-sulfate----iduronic acid----N-acetylglucosamine 6-O-sulfate----glucuronic acid----N-sulfated glucosamine 3,6-di-O-sulfate----iduronic acid 2-O-sulfate----anhydromannitol 6-O-sulfate, is resistant to degradation by both enzymes and binds antithrombin with a dissociation constant of (4-18) X 10(-7) M. The occurrence of a 15-17% replacement of N-sulfated glucosamine 3,6-di-O-sulfate with N-sulfated glucosamine 3-O-sulfate and a 10-12% replacement of iduronic acid with glucuronic acid in both octasaccharides indicates that these substitutions have little or no effect on the binding of the oligosaccharides to the protease inhibitor. When bound to antithrombin, both octasaccharides produce a 40% enhancement in the intrinsic fluorescence of the protease inhibitor and a rate of human factor Xa inhibition of 5 X 10(5) M-1 s-1 as monitored by stopped-flow fluorometry. This suggests that the conformation of antithrombin in the region of the factor Xa binding site is similar when the protease inhibitor is complexed with either octasaccharide.  相似文献   

18.
A heparin-binding peptide within antithrombin III (ATIII) was identified by digestion of ATIII with Staphylococcus aureus V8 protease followed by purification on reverse-phase high pressure liquid chromatography using a C-4 column matrix. The column fractions were assayed for their ability to bind heparin by ligand blotting with 125I-fluoresceinamine-heparin as previously described (Smith, J. W., and Knauer, D. J. (1987) Anal. Biochem. 160, 105-114). This analysis identified at least three fractions with heparin binding ability of which the peptide eluting at 25.4 min gave the strongest signal. Amino acid sequence analysis of this peptide gave a partially split sequence which was consistent with regions encompassing amino acids 89-96 and 114-156. These amino acids are present in a 1:1 molar ratio which is consistent with a disulfide linkage between Cys-95 and Cys-128. High affinity heparin competed more effectively for the binding of 125I-fluoresceinamine-heparin to this peptide than low affinity heparin. Chondroitin sulfate did not block the binding of 125I-fluoresceinamine-heparin to the peptide. These data strongly suggest that the isolated peptide represents a native heparin-binding region within intact ATIII. Computer generation of a plot of running charge density of ATIII confirms that the region encompassing amino acid residues 123-141 has the highest positive charge density within the molecule. A hydropathy plot of ATIII was generated using a method similar to that of Kyte and Doolittle (Kyte, J., and Doolittle, R. F. (1982) J. Mol. Biol. 157, 105-132). This plot indicates that amino acid residues 126-140 are exposed to the exterior surface of the molecule. Based on these data, we suggest that the region corresponding to amino acid residues 114-156 is a likely site for the physiological heparin-binding domain of ATIII. We also conclude that the proposed disulfide bridges within the protein are suspect and should be re-examined (Petersen, T. E., Dudek-Wojiechowska, G., Sottrup-Jensen, L., and Magnussun, S. (1979) in The Physiological Inhibitors of Coagulation and Fibrinolysis (Collen, D., Wiman, B., and Verstaeta, M., eds) pp. 43-54, Elsevier Scientific Publishing Co., Amsterdam).  相似文献   

19.
In this work, we describe a method of constructing a film of linear poly(glycidyl methacrylate) (PGMA) polymer onto the surface of quartz crystal microbalance (QCM) electrode as a coating material that allows easy coupling of heparin molecules onto the electrode and facilitates the determination of the interaction between heparin and antithrombin III (AT III). The PGMA film was characterized with atomic force microscopy (AFM) and infra-red spectroscopy. The coupling of heparin was accomplished in one step solution reaction. A home-made quartz crystal microbalance-flow injection analysis (QCM-FIA) system with data analysis software developed in our laboratory was used to determine the interaction. The interactions between immobilized heparin and AT III were studied with various concentrations under various conditions. The obtained constants are kass=(1.49+/-0.12)x10(3)mol-1ls-1, kdiss=(3.94+/-0.63)x10(-2)s-1, KA=(3.82+/-0.33)x10(4)mol-1l.  相似文献   

20.
The LexA repressor from Escherichia coli is a sequence-specific DNA binding protein that shows no pronounced sequence homology with any of the known structural motifs involved in DNA binding. Since little is known about how this protein interacts with DNA, we have selected and characterized a great number of intragenic, second-site mutations which restored at least partially the activity of LexA mutant repressors deficient in DNA binding. In 47 cases, the suppressor effect of these mutations was due to an Ind- phenotype leading presumably to a stabilization of the mutant protein. With one exception, these second-site mutations are all found in a small cluster (amino acid residues 80 to 85) including the LexA cleavage site between amino acid residues 84 and 85 and include both already known Ind- mutations as well as new variants like GN80, GS80, VL82 and AV84. The remaining 26 independently isolated second-site suppressor mutations all mapped within the amino-terminal DNA binding domain of LexA, at positions 22 (situated in the turn between helix 1 and helix 2) and positions 57, 59, 62, 71 and 73. These latter amino acid residues are all found beyond helix 3, in a region where we have previously identified a cluster of LexA (Def) mutant repressors. In several cases the parental LexA (Def) mutation has been removed by subcloning or site-directed mutagenesis. With one exception, these LexA variants show tighter in vivo repression than the LexA wild-type repressor. The most strongly improved variant (LexA EK71, i.e. Glu71----Lys) that shows an about threefold increased repression rate in vivo, was purified and its binding to a short consensus operator DNA fragment studied using a modified nitrocellulose filter binding assay. As expected from the in vivo data, LexA EK71 interacts more tightly with both operator and (more dramatically) with non-operator DNA. A determination of the equilibrium association constants of LexA EK71 and LexA wild-type as a function of monovalent salt concentration suggests that LexA EK71 might form an additional ionic interaction with operator DNA as compared to the LexA wild-type repressor. A comparison of the binding of LexA to a non-operator DNA fragment further shows that LexA interacts with the consensus operator very selectively with a specificity factor of Ks/Kns of 1.4 x 10(6) under near-physiological salt conditions.  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