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.
Thrombin-inhibitory activity of whale heparin oligosaccharides   总被引:1,自引:0,他引:1  
Whale heparin was partially digested with a purified heparinase and the oligosaccharide fractions with 8-20 monosaccharide units were isolated from the digest by gel filtration on Sephadex G-50, followed by affinity chromatography on a column of antithrombin III immobilized on Sepharose 4B. A marked difference in the inhibitory activity for thrombin in the presence of antithrombin III was observed between the high-affinity fractions for antithrombin III of octasaccharide approximately hexadecasaccharide and those of octadecasaccharide approximately eicosasaccharide. The disaccharide compositions of these hexadeca-, octadeca-, and eicosasaccharides were analyzed by high-performance liquid chromatography after digestion with a mixture of purified heparitinases 1 and 2 and heparinase. The analytical data indicated that the proportions of trisulfated disaccharide (IdUA(2S)alpha 1----4GlcNS(6S)) and disulfated disaccharide (UA1----4GlcNS(6S)) increased with the manifestation of high thrombin-inhibitory activity, while that of monosulfated disaccharide (UA1----4GlcNS) decreased. The present observations, together with those so far reported, suggest that the presence of the former structural elements, specifically IdUA(2S)alpha 1----4GlcNS(6S), as well as the antithrombin III-binding pentasaccharide at the proper positions in the molecules of whale heparin oligosaccharides is essential for the manifestation of high inhibitory activity for thrombin in the presence of antithrombin III. The structural bases for the manifestation of the anticoagulant activity of whale and porcine heparins and their oligosaccharides are also discussed.  相似文献   

2.
Porcine heparin, whale heparin, and a solvolyzed porcine heparin were deaminated, and sulfated oligosaccharides, compounds 3f, 4f, 3s, 4s, 5, 6, 7s, 10, 11f, 11s, and 13 were isolated from the deamination products by Dowex 1 x 2 (Cl- form) column chromatography and high voltage paper electrophoresis and/or gel filtration on Sephadex G-25. Based on the results of chemical, 1H and 13C NMR spectral analyses, and of Smith degradation, together with previous observations, the structures of these sulfated oligosaccharides are proposed to be as follows: compound 3f, IdUA(2S)alpha 1 leads to GlcNAc alpha 1 leads to 4GlcUA; compound 4f, IdUA alpha 1 leads to 4GlcNAc(6S) alpha 1 leads to 4GlcUA; compound 3s, IdUA(2S) alpha 1 leads to 4GlcNAc alpha 1 leads to 4 GlcUA beta 1 leads to 4a Man; compound 4s, IdUA alpha 1 leads to 4Glc NAc(6S) alpha 1 leads to 4 GlcUA beta 1 leads to 4aMan; compound 5, IdUA(2S) alpha 1 leads to 4aMan; compound 6, GlcUA beta 1 leads to aMan(6S); compound 7s, IdUA alpha 1 leads to 4aMan(6S); compound 10, IdUA(2S)alpha 1 leads 4GlcNAc(6S)alpha 1 leads to 4 GlcUA beta 1 leads to 4aMan; compound 11f, IdUA(2S) alpha 1 leads 4GlcNAc alpha 1 leads to 4GlcUA beta 1 leads to 4a Man (6S); compound 11s, IdUA alpha 1 leads to GlcNAc(6S) alpha 1 leads to 4GlcUA beta 1 leads to 4aMan(6S); compound 13, IdUA(2S) alpha 1 leads to 4aMan(6S). For ths sulfated disaccharides, the same results as those reported in our previous papers were obtained. On the other hand, the proportion of total sulfated tri- and tetrasaccharides from whale heparin was 1.9 times higher than that from porcine heparine, reflecting a higher content of GlcNAc in the former. Also, the yields of compound 11s from these two heparins were comparable to their anticoagulant activities. In addition, certain 2-O-sulfates on IdUA flanked with GlcNS(6X) (X=H or S) in the heparin molecule are suggested to be important for the activity.  相似文献   

3.
The interference of the heparin-neutralizing plasma component S protein (vitronectin) (Mr = 78,000) with heparin-catalyzed inhibition of coagulation factor Xa by antithrombin III was investigated in plasma and in a purified system. In plasma, S protein effectively counteracted the anticoagulant activity of heparin, since factor Xa inhibition was markedly reduced in comparison to heparinized plasma deficient in S protein. Using purified components in the presence of heparin, S protein induced a concentration-dependent reduction of the inhibition rate of factor Xa by antithrombin III. This resulted in a decrease of the apparent pseudo-first order rate constant by more than 10-fold at a physiological ratio of antithrombin III to S protein. S protein not only counteracted the anticoagulant activity of commercial heparin but also of low molecular weight forms of heparin (mean Mr of 4,500). The heparin-neutralizing activity of S protein was found to be mainly expressed in the range 0.2-10 micrograms/ml of high Mr as well as low Mr heparin. S protein and high affinity heparin reacted with apparent 1:1 stoichiometry to form a complex with a dissociation constant KD = 1 X 10(-8) M as determined by a functional assay. As deduced from dot-blot analysis, direct interaction of radiolabeled heparin with S protein revealed a dissociation constant KD = 4 X 10(-8) M. Heparin binding as well as heparin neutralization by S protein increased significantly when reduced/carboxymethylated or guanidine-treated S protein was employed indicating the existence of a partly buried heparin-binding domain in native S protein. Radiolabeled heparin bound to the native protein molecule as well as to a BrCN fragment (Mr = 12,000) containing the heparin-binding domain as demonstrated by direct binding on nitrocellulose replicas of sodium dodecyl sulfate-polyacrylamide gels. Kinetic analysis revealed that the heparin neutralization activity of S protein in the inhibition of factor Xa by antithrombin III could be mimicked by a synthetic tridecapeptide from the amino-terminal portion of the heparin-binding domain. These data provide evidence that the heparin-binding domain of S protein appears to be unique in binding to heparin and thereby neutralizing its anticoagulant activity in the inhibition of coagulation factors by antithrombin III. The induction of heparin binding and neutralization may be considered a possible physiological mechanism initiated by conformational alteration of the S protein molecule.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

4.
The rate of the reaction between thrombin and antithrombin III is greatly increased in the presence of heparin. Several mechanisms for this effect are possible. To study the problems commercial heparin was fractionated into one fraction of high anticogulant activity and one of low anticoagulant activity by affinity chromatography on matrix-bound antithrombin III. The strength of the binding of the two heparin fractions to antithrombin III and thrombin, respectively, was determined by a crossed immunoelectrophoresis technique. As was to be expected, the high activity fraction was strongly bound to antithrombin III while the low activity fraction was weakly bound. In contrast, thrombin showed equal binding affinity for both heparin fractions. The ability of the two heparin fractions to catalyse the inhibition of thrombin by antithrombin III was determined and was found to be much greater for the high activity heparin fraction. A mechanism for the reaction between thrombin and antithrombin III in the presence of small amounts of heparin is suggested, whereby antithrombin III first binds heparin and this complex then inhibits thrombin by interaction with both the bound heparin and the antithrombin III.  相似文献   

5.
Monoclonal antibodies were raised against human thrombin-antithrombin III complex by a hybridoma technique. Among them, five monoclonal antibodies, designated as JITAT-4, -14, -16, -17 and -19, were found to react with thrombin-antithrombin III, but not with its nascent components, α-thrombin or antithrombin III. Their respective immunoglobulin classes are IgG1 for JITAT-16 and -19, and IgG2a for JITAT-4, -14 and -17. Besides the thrombin-antithrombin III complex, they all bound to the Factor Xa-antithrombin III complex and the active-site-cleaved two-chain antithrombin III as well. Moreover, the reactivity of these two antibodies to the neoantigens was not affected by heparin, suggesting that their epitopes are independent of heparin-induced conformational changes of antithrombin III. Two of them, JITAT-16 and -17, were categorized as high-affinity antibodies to thrombin-antithrombin III complex, the dissociation constants being 6.7 nM and 4.8 nM, respectively. However, they do not share antigenic determinants. These monoclonal antibodies may allow us to explore more precisely the reaction between antithrombin III and thrombin or its related enzymes.  相似文献   

6.
Human neutrophil elastase catalyzes the inactivation of antithrombin by a specific and limited proteinolytic cleavage. This inactivation reaction is greatly accelerated by an active anticoagulant heparin subfraction with high binding affinity for antithrombin. A potentially complex reaction mechanism is suggested by the binding of both neutrophil elastase and antithrombin to heparin. The in vitro kinetic behavior of this system was examined under two different conditions: 1) at a constant antithrombin concentration in which the active anticoagulant heparin was varied from catalytic to saturating levels; and 2) at a fixed, saturating heparin concentration and variable antithrombin levels. Under conditions of excess heparin, the inactivation could be continuously monitored by a decrease in the ultraviolet fluorescence emission of the inhibitor. A Km of approximately 1 microM for the heparin-antithrombin complex and a turnover number of approximately 200/min was estimated from these analyses. Maximum acceleratory effects of heparin on the inactivation of antithrombin occur at heparin concentrations significantly lower than those required to saturate antithrombin. The divergence in acceleratory effect and antithrombin binding contrasts with the anticoagulant functioning of heparin in promoting the formation of covalent antithrombin-enzyme complexes and is likely to derive from the fact that neutrophil elastase is not consumed in the inactivation reaction. A size dependence was observed for the heparin effect since an anticoagulantly active octasaccharide fragment of heparin, with avid antithrombin binding activity, was without effect on the inactivation of antithrombin by neutrophil elastase. Despite the completely nonfunctional nature of elastase-cleaved antithrombin and the altered physical properties of the inhibitor as indicated by fluorescence and sodium dodecyl sulfate-polyacrylamide gel electrophoresis, the inactivated inhibitor exhibited a circulating half-life in rabbits that was indistinguishable from native antithrombin. These results point to an unexpected and apparently contradictory function for heparin which may relate to the properties of the vascular endothelium in pathological situations.  相似文献   

7.
A Horne  P Gettins 《Biochemistry》1992,31(8):2286-2294
The effects of length and composition upon the antithrombin-binding properties of heparin have been investigated for two series of structurally related heparin oligosaccharides. Each series consists of a tetrasaccharide, hexasaccharide, and octasaccharide heparin fragment composed of alternating hexuronic acid (either iduronate 2-sulfate or glucuronate) and glucosamine 6,N-disulfate residues. These two series represent dominant structural motifs in intact heparin and differ from each other by the presence of a glucuronic acid in one series in place of an iduronate 2-sulfate residue penultimate to the reducing end of the fragment. Perturbations to the 1H resonances in the NMR spectrum of antithrombin upon binding of the two series of heparin fragments are compared to those generated by intact heparin binding, as well as to the effects of binding of a synthetic high-affinity pentasaccharide. All of the heparin fragments examined appear to bind to antithrombin at the same site. Three of the heparin fragments (hexasaccharide-2, octasaccharide-2, and octasaccharide-1) produce almost identical perturbations in the antithrombin 1H NMR spectrum compared to binding of intact heparin, including perturbations of resonances from tryptophan 49. This indicates that neither the glucuronic acid nor the trisulfated glucosamine residue (structural elements known to be part of the high-affinity heparin motif) are necessary for the majority of the conformational changes induced upon heparin fragment binding to antithrombin. However, the low anticoagulant activity of these fragments indicates that the changes in protein conformation upon fragment binding, as manifested by these 1H resonance perturbations, are not sufficient for catalytic activation of the inhibitor.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
S protein, a major inhibitor of the assembly of the membrane attack complex of complement, has recently been shown to be identical to the serum spreading factor vitronectin. It also neutralizes the anticoagulant activities of heparin. We have studied the structural requirements for the heparin neutralizing properties of S protein/vitronectin using heparin, heparan sulfate, and heparin oligosaccharides with well defined anticoagulant specificities. The abilities of heparin fractions, Mr 7,800-18,800, with high affinity for antithrombin, and of the International Heparin Standard, to accelerate the inactivation of thrombin and Factor Xa by antithrombin were readily neutralized by S protein/vitronectin. Binding and neutralization of heparin by S protein/vitronectin was inhibited by heparin with low affinity for antithrombin, indicating that S protein/vitronectin can interact with a region on the heparin chain that might serve as a proteinase binding site. S protein/vitronectin efficiently neutralized oligosaccharides of Mr 2,400-7,200, unlike the two other physiologically occurring heparin neutralizing proteins histidine-rich glycoprotein and platelet factor 4. Furthermore, S protein/vitronectin neutralized the anti-Factor Xa activity of a synthetic pentasaccharide comprising the antithrombin-binding sequence of heparin. High molar excess of a synthetic tridecapeptide corresponding to part (amino acids 374-359) of the proposed glycosaminoglycan binding domain of S protein/vitronectin neutralized high affinity heparin and some oligosaccharides, but failed to neutralize the synthetic antithrombin-binding pentasaccharide. Like platelet factor 4, but unlike histidine-rich glycoprotein, S protein/vitronectin readily neutralized the anticoagulant activities of heparan sulfate of Mr approximately 20,000. These findings suggest that S protein/vitronectin may interact through its glycosaminoglycan binding domain(s) with various functional domains of the heparin (heparan sulfate) molecule, including the antithrombin-binding pentasaccharide sequence. Furthermore, the results suggest that S protein/vitronectin may be a physiologically important modulator of the anticoagulant activity of heparin-like material on or near the vascular endothelium.  相似文献   

9.
Cultured microvascular endothelial cells isolated from rat epididymal fat pads produce glycosaminoglycans that accelerate thrombin-antithrombin complex formation. The heparinlike nature of these macromolecules was established by complete destruction of their anticoagulant activity employing purified Flavobacterium heparinase. Only 15% of the biologic activity of these complex carbohydrates was expressed when the heparin binding domain on the protease inhibitor was chemically modified at the Trp 49 residue. The anticoagulantly active species contains disaccharides which constitute the unique antithrombin binding region of the mucopolysaccharide. Removal of the biologically active heparinlike components from endothelial cells with 0.05% trypsin suggests that these molecular species are present on the cell surface.  相似文献   

10.
To map out the heavy metal binding sites of iduronic acid containing oligosaccharides isolated from human kidneys, we studied Zn(II) binding by nuclear magnetic resonance (NMR) and molecular modeling to two disaccharides isolated after nitrous acid depolymerization of heparin and two synthetic disaccharides representative of the heparin structure, namely, IdopA2S (alpha 1,4)AnManOH, 1 alpha, IdopA2S (alpha 1,4)AnManOH6S, 1b, IdopA2S-(alpha 1,4)GlcNS alpha Me, 2a, and IdopA2S (alpha 1,4)GlcNS6S alpha Me, 2b (see previous article in this series). A conformational analysis of the metal free and metal bound solutions was made by comparing calculated [(NOE)]s, [T1]s, and [J]s to experimental values. The 1C4, 4C1, and 2S0 conformations of the L-idopyranosiduronate ring and the 4E and 4T3 of the anhydro-D-mannitol ring are evaluated as are rotations about the C5-C6 hydroxymethylene of the AnManOH(6S) or GlcNS (6S) residues. The NOE between IdopA2S H1 and H3 and the known NOE between H2 and H5, as well as the T1 of IdopA2S H3, are introduced as NMR observables sensitive to the IdopA2S ring conformation. Similarly, a NOE between IdopA2S H5 and AnManOH(6S) or GlcNS(6S) H3 was observed that directly restricts the allowed interglycosidic conformational space. For all disaccharides, the Zn(II) bound spectral data are consistent with models in which these motions are partially "frozen" such that the 1C4 conformation of the IdopA2S is stabilized along with the 4T3 conformation of the AnManOH(6S) ring. The interglycosidic conformation is also stabilized in one of two minima. Electrostatic potential energy calculations gave the best overall agreement with experiment and suggest metal binding conformations with the carboxylate and ring oxygen of the IdopA2S residues (1C4 conformation) and either O3 of the GlcNS(6S) residues or the sulfate oxygens of the 6-sulphate for 2b providing additional chelating sites. These chelation models concur with the observation of marked 13C and 1H NMR chemical shifts for the IdopA2S resonances and of GlcNS H3 for 2 alpha and GlcNS6S C6 for 2b. This study of model compounds implicates the IdopA2S(alpha 1,4)GlcNS6S group as part of the heavy metal binding site in biologically important acidic oligosaccharides such as heparin.  相似文献   

11.
The specificity of endothelial binding sites for heparin was investigated with heparin fractions and fragments differing in their Mr, charge density and affinity for antithrombin III, as well as with heparinoids and other anionic polyelectrolytes (polystyrene sulphonates). The affinity for endothelial cells was estimated by determining I50 values in competition experiments with 125I-heparin. We found that affinity for endothelial cells increases as a function of Mr and charge density (degree of sulphation). Binding sites are not specific receptors for heparin. Other anionic polyelectrolytes, such as pentosan polysulphates and polystyrene sulphonates, competed with heparin for binding to endothelial cells. Fractions of standard heparin with high affinity for antithrombin III also had greater affinity for endothelium. However, these two properties of heparin (affinity for antithrombin III and affinity for endothelial cells) could be dissociated. Oversulphated heparins and oversulphated low-Mr heparin fragments had lower anticoagulant activity and higher affinity for endothelial cells than did their parent compounds. Synthetic pentasaccharides, bearing the minimal sequence for binding to antithrombin III, did not bind to endothelial cells. Binding to endothelial cells involved partial neutralization of heparin. Bound heparin exhibited only 5% and 7% of antifactor IIa and antifactor Xa specific activity, respectively. In the presence of 200 nM-antithrombin III, and in the absence of free heparin, a limited fraction (approx. 30%) of bound heparin was displaced from endothelial cells during a 1 h incubation period. These data suggested that a fraction of surface-bound heparin could represent a pool of anticoagulant.  相似文献   

12.
The tributylammonium salt of whale (Balaenoptera borealis L.) intestinal heparin with high affinity for antithrombin III, whose degrees of sulfate-substitution in D-glucosamine and L-iduronic acid residues are GlcNS 0.738, GlcN6S 0.384, and IdoA2S 0.510 mol, was reacted with 2.5, 5.0, or 10.0 mol of pyridine-sulfur trioxide/mol of available hydroxyl groups in N,N-dimethylformamide at -10 degrees C for 1 h. Both chemical and NMR spectroscopic analyses revealed that an exclusive 6-O-sulfation of the D-glucosamine residues proceeded, according to the amount of the sulfating reagent used (GlcN6S: 0.476, 0.585, and 0.641 mol, respectively), the degree of sulfation at other natural substitution positions in the polysaccharide being unchanged, without any detectable unnatural sulfate-substitution. Biological examination of these products indicated that the 6-O-sulfation in the original whale heparin resulted in significant increases in blood clotting and anti-Factor IIa activities (maximal 43 and 82% increases, respectively), and in a moderate increase in the ability to bind antithrombin III, that is, in anti-Factor Xa activity and in intrinsic fluorescence enhancement of the protein (maximal 28 and 30% increases, respectively), together with a maximal 10% increase in the proportion of heparin species with higher affinity for antithrombin III, released with 1.0-3.0 M NaCl from antithrombin III-Sepharose.  相似文献   

13.
Adult male rats were given [35S]sulphate intraperitoneally. Heparan [35S]sulphate (HS) chains were recovered from adipose tissue, brain, carcase, heart, intestine, kidneys, liver, lungs, skin and spleen by digestion with Pronase, precipitation with cetylpyridinium chloride, digestion with chondroitin ABC lyase and DNAase and gradient elution from DEAE-Sephacel. Purity was confirmed by agarose-gel electrophoresis and degradation with HNO2. Fractionation by gradient elution from antithrombin-agarose indicated that the proportion of HS with high binding affinity for antithrombin (HA-HS) ranged from 4.7% (kidneys) to 21.5% (brain). On a mass basis the major sources of HA-HS were carcase, skin and intestine. HA-HS from intestine was arbitrarily divided into subfractions I-VI, with anticoagulant activities ranging from 1 to 60 units/mg [by amidolytic anti-(Factor IIa) assay] and from 4 to 98 units/mg [by amidolytic anti-(Factor Xa) assay], indicating that the antithrombin-binding-site densities of HA-HS chains covered a wide range, as shown previously for rat HA-heparin chains [Horner, Kusche, Lindahl & Peterson (1988) Biochem. J. 251, 141-145]. HA-HS subfractions II, IV and VI were mixed with samples of HA-[3H]heparin chains and rechromatographed on antithrombin-agarose. Affinity for matrix-bound antithrombin did not correlate with anticoagulant activity, e.g. HA-HS subfraction IV [38 anti-(Factor Xa) units/mg] was co-eluted with HA-heparin chains [127 anti-(Factor Xa) units/mg].  相似文献   

14.
Heparin and heparin oligosaccharides prepared by nitrous acid depolymerization were fractionated by affinity chromatography on immobilized antithrombin and by gel chromatography. The anticoagulant activities of high affinity heparin of Mr greater than or equal to 7,800 could be readily neutralized by the plasma protein histidine-rich glycoprotein (see also Lijnen, H.R., Hoylaerts, M., and Collen, D. (1983) J. Biol. Chem. 258, 3803-3808), whereas oligosaccharides falling below 18 saccharide units (Mr 5,400) became increasingly resistant to neutralization. An octasaccharide with characteristic marked ability to accelerate the inactivation of Factor Xa by antithrombin retained greater than 50% of its activity even at a histidine-rich glycoprotein/oligosaccharide molar ratio of 500:1. Histidine-rich glycoprotein, like the platelet-derived heparin neutralizing protein platelet factor 4 (Lane, D.A., Denton, J., Flynn, A.M., Thunberg, L. and Lindahl, U. (1984) Biochem J. 218, 725-732), therefore requires interaction with saccharide sequences in addition to the antithrombin-binding pentasaccharide of heparin in order to efficiently express its antiheparin activity. Heparan sulfate isolated from pig intestinal mucosa (HS I, Mr approximately 20,000) and from human aorta (HS II, Mr approximately 40,000) exhibited anti-Factor Xa activities of 180 and 20 units/micromol [corrected], respectively. A fraction corresponding to about 5% of HS I bound with high affinity to immobilized antithrombin and contained all of the anticoagulant activity of the starting material. While these heparan sulfates were readily neutralized by platelet factor 4, they were relatively resistant to neutralization by histidine-rich glycoprotein, although complete neutralization could be attained in the presence of molar excess of this protein. These findings may be of importance in relation (a) to the functional role of endogenous anticoagulant polysaccharides at the vascular wall and (b) to clinical situations in which heparin or heparin-related compounds are administered as exogenous anticoagulants.  相似文献   

15.
The isolation, some structural features, physicochemical properties and pharmacological activities of a heparin from Anomalocardia brasiliana are reported. It is shown that the mollusc heparin is very similar to those present in mammalian tissues with regard to chemical composition, physicochemical properties, pharmacological activities and susceptibility to heparinase and heparitinase II from Flavobacterium heparinum, as well as to the types of products formed by the action of these enzymes. Three significant quantitative differences were observed for the mollusc heparin when compared with the ones from mammalian origin, namely, a higher degree of binding with antithrombin III (45%), higher molecular weight (27-43 kDa) and higher anticoagulant activity (320 I.U./mg). The possible biological role of heparin is discussed in view of the present findings.  相似文献   

16.
Inhibition of bovine factor IXa and factor Xabeta by antithrombin III.   总被引:10,自引:0,他引:10  
Factor IXa and factor Xabeta are serine proteases which participate in the middle phase of blood coagulation. These two enzymes are inhibited by antithrombin III by the formation of an enzyme-inhibitor complex containing 1 mol of enzyme and 1 mol of antithrombin III. The complex was readily demonstrated by sodium dodecyl sulfate polyacrylamide gel electrophoresis and loss of coagulant or esterase activity at increasing concentrations of inhibitor. The inactivation of factor IXa by antithrombin III was relatively slow, but the reaction was greatly accelerated by the addition of heparin.  相似文献   

17.
The effect of various well-characterized heparin preparations on the inactivation of human Factor XIa by human antithrombin III was studied. The heparin preparations used were unfractionated heparin and four heparin fractions obtained after anion-exchange chromatography. Inactivation of Factor XIa was monitored with S2366 as chromogenic substrate and followed pseudo-first-order reaction kinetics under all reaction conditions tested. Enhancement of the rate of inhibition of Factor XIa in the presence of unfractionated heparin correlated to the binding of antithrombin III to heparin. From the kinetic data a binding constant of 0.1 microM was inferred. The maximum rate enhancement, achieved at saturating heparin concentrations, was 30-fold. The rate enhancement achieved in the presence of each of the heparin fractions could also be correlated to the binding of antithrombin III to the heparin. The binding constant inferred from the kinetic data varied from 0.10 to 0.28 microM and the number of binding sites for antithrombin III varied from 0.06 to 0.74 site per heparin molecule. The maximum rate enhancements, achieved at saturating heparin concentrations, were strongly dependent on the type of heparin used and varied from 7-fold for fraction A to 41-fold for fraction D. Therefore, although the stimulation of Factor XIa inactivation by antithrombin III could be quantitatively correlated to the binding of antithrombin III to heparin, the heparin-catalysed inhibition of Factor XIa is dependent not only upon the degree of binding of antithrombin III to heparin but also upon the type of heparin to which antithrombin III is bound.  相似文献   

18.
Fucoidan, poly(L-fucopyranose) linked primarily alpha 1----2 with either a C3- or a C4-sulfate, is an effective anticoagulant in vitro and in vivo (Springer, G. F., Wurzel, H. A., McNeal, G. M., Jr., Ansell, N. J., and Doughty, M. F. (1957) Proc. Soc. Exp. Biol. Med. 94, 404-409). We have determined the antithrombin effects of fucoidan on the glycosaminoglycan-binding plasma proteinase inhibitors antithrombin III and heparin cofactor II. Fucoidan enhances the heparin cofactor II-thrombin reaction more than 3500-fold. The apparent second-order rate constant of thrombin inhibition by heparin cofactor II increases from 4 x 10(4) (in the absence of fucoidan) to 1.5 x 10(8) M-1 min-1 as the fucoidan concentration increases from 0.1 to 10 micrograms/ml and then decreases as fucoidan is increased above 10 micrograms/ml. The fucoidan reaction with heparin cofactor II-thrombin is kinetically equivalent to a "template model." Apparent fucoidan-heparin cofactor II and fucoidan-thrombin dissociation constants are 370 and 1 nM, respectively. The enhancement of thrombin inhibition by fucoidan, like heparin and dermatan sulfate, is eliminated by selective chemical modification of lysyl residues either of heparin cofactor II or of thrombin. The fucoidan-antithrombin III reactions with thrombin and factor Xa are accelerated maximally 285- and 35-fold at fucoidan concentrations of 30 and 500 micrograms/ml, respectively. Using human plasma and 125I-labeled thrombin in an ex vivo system, the heparin cofactor II-thrombin complex is formed preferentially over the antithrombin III-thrombin complex in the presence of 10 micrograms/ml fucoidan. Our results indicate that heparin cofactor II is activated by fucoidan in vitro and in an ex vivo plasma system and suggest that the major antithrombin activity of fucoidan in vivo is mediated by heparin cofactor II and not by antithrombin III.  相似文献   

19.
Tryptase from human mast cells has been shown (in vitro) to catalyze the destruction of fibrinogen and high-molecular-weight kininogen as well as the activation of C3a and collagenase. Although large amounts of tryptase are released in tissues by degranulating mast cells and levels as high as 1000 ng/ml have been measured in the circulation following systemic anaphylaxis, no specific physiologic inhibitor has yet been found for the protease. The current work tests several more inhibitors for their effects on tryptase and examines any effect of tryptase on these inhibitors. First, antileukoprotease and low-molecular-weight elastase inhibitor from human lung and hirudin and antithrombin III had no effect on tryptase activity in vitro. Second, the possibility that tryptase, being insensitive to the effects of inhibitors, might instead destroy them was also considered. Tryptase failed to cleave and inactivate antileukoprotease, low-molecular-weight elastase inhibitor, alpha 1 protease inhibitor, alpha 2 macroglobulin, and antithrombin III. Third, based on the knowledge that tryptase stability is regulated by its interaction with heparin, antithrombin III was used as a model heparin-binding protein to demonstrate that a protein competitor for heparin-binding sites, presumably by displacement of tryptase, destabilizes this enzyme. Conversely, tryptase, in excess, blocked the binding of antithrombin III to heparin, thereby attenuating the heparin-mediated inhibition of thrombin by antithrombin III.  相似文献   

20.
Oligosaccharides of well-defined molecular size were prepared from heparin by nitrous acid depolymerization, affinity chromatography on immobilized antithrombin III (see footnote on Nomenclature) and gel chromatography on Sephadex G-50. High affinity (for antithrombin III) octa-, deca-, dodeca-, tetradeca-, hexadeca- and octadeca-saccharides were prepared, as well as oligosaccharides of larger size than octadecasaccharide. The inhibition of Factor Xa by antithrombin III was greatly accelerated by all of these oligosaccharides, the specific anti-Factor Xa activity being invariably greater than 1300 units/mumol. The anti-Factor Xa activity of the decasaccharide was not significantly decreased in the presence of platelet factor 4, even at high platelet factor 4/oligosaccharide ratios. Measurable but incomplete neutralization of the anti-Factor Xa activities of the tetradeca- and hexadeca-saccharides was observed, and complete neutralization of octadeca- and larger oligo-saccharides was achieved with excess platelet factor 4. The octa-, deca-, dodeca-, tetradeca- and hexadeca-saccharides had negligible effect on the inhibition of thrombin by antithrombin III, whereas specific anti-thrombin activity was expressed by the octadeca-saccharide and by the larger oligosaccharides. An octadecasaccharide is therefore the smallest heparin fragment (prepared by nitrous acid depolymerization) that can accelerate thrombin inhibition by antithrombin III. The anti-thrombin activities of the octadecasaccharide and larger oligosaccharides were more readily neutralized by platelet factor 4 than were their anti-Factor Xa activities. These findings are compatible with two alternative mechanisms for the action of platelet factor 4, both involving the binding of the protein molecule adjacent to the antithrombin III-binding site. Such binding results in either steric interference with the formation of antithrombin III-proteinase complexes or in displacement of the antithrombin III molecule from the heparin chain.  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