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.
A heparin preparation obtained by gel chromatography is compared to unfractionated heparin with respect to the effects of heparin on the reaction between thrombin and heparin cofactor. Whereas both preparations enhance the rate of inhibition of thrombin by heparin cofactor, the extent of inhibition is decreased by the unfractionated, but not by the fractionated heparin. The decreased extent of inhibition is accounted for by residua of unreacted and undegraded heparin cofactor and thrombin, as demonstrated by gel electrophoresis in dodecyl sulfate. However both heparin preparations enhance the rate of degradation by thrombin of the thrombin-heparin cofactor complex.  相似文献   

2.
Only 30% of commercial heparin reacts with antithrombin III (ATIII). This study shows that the interaction is pH dependent: 100% of the heparin binds to ATIII at pH 3.0, 30% at physiological pH. Binding of ATIII, platelet factor 4, and protamine to heparin was studied using a new rocket precipitin electrophoresis procedure, adapted from the Laurell rocket immunoelectrophoresis procedure. Protamine is incorporated into agarose gel, and heparin mixtures with protamine, ATIII, or platelet factor 4 electrophoresed into the gel from a series of wells. The residual free heparin is precipitated by the protamine in a rocket-shaped arc, the height of which is proportional to the amount of free heparin. No antibody is employed. This procedure is useful for quantitation of heparin and for studying the binding of heparin to proteins.  相似文献   

3.
Activation of heparin cofactor II by heparin oligosaccharides   总被引:1,自引:0,他引:1  
Heparin was partially depolymerized with heparinase or nitrous acid. The resulting oligosaccharides were fractionated by gel filtration chromatography and tested for the ability to stimulate inhibition of thrombin by purified heparin cofactor II or antithrombin. Oligosaccharides containing greater than or equal to 18 monosaccharide units were active with antithrombin, while larger oligosaccharides were required for activity with heparin cofactor II. Intact heparin molecules fractionated on a column of immobilized antithrombin were also tested for activity with both inhibitors. The relative specific activities of the unbound heparin molecules were 0.06 with antithrombin and 0.76 with heparin cofactor II in comparison to unfractionated heparin (specific activity = 1.00). We conclude that heparin molecules much greater than 18 monosaccharide units in length are required for activity with heparin cofactor II and that the high-affinity antithrombin-binding structure of heparin is not required.  相似文献   

4.
The binding of heparin causes a conformational change in antithrombin to give an increased heparin binding affinity and activate the inhibition of thrombin and factor Xa. The areas of antithrombin involved in binding heparin and stabilizing the interaction in the high-affinity form have been partially resolved through the study of both recombinant and natural variants. The role of a section of the N-terminal segment of antithrombin, residues 22-46 (segment 22-46), in heparin binding was investigated using rapid kinetic analysis of the protein cleaved at residues 29-30 by limited proteolysis with thermolysin. The cleaved antithrombin had 5.5-fold lowered affinity for heparin pentasaccharide and 1.8-fold for full-length, high-affinity heparin. It was shown that, although the initial binding of heparin is slightly enhanced by the cleavage, it dissociates much faster from the cleaved form, giving rise to the overall decrease in heparin affinity. This implies that the segment constituting residues 22-46 in the N terminus of antithrombin hinders access to the binding site for heparin, hence the increased initial binding for the cleaved form, whereas, when heparin is bound, segment 22-46 is involved in the stabilization of the binding interaction, as indicated by the increased dissociation constant. When the heparin pentasaccharide is bound to antithrombin prior to incubation with thermolysin, it protects the N-terminal cleavage site, implying that segment 22-46 moves to interact with heparin in the conformational change and thus stabilizes the complex.  相似文献   

5.
Previous reports have shown that heparin is an inhibitor of casein kinase-2 (CK-2). It is unclear whether heparin is also an inhibitor of glycogen synthase (casein) kinase-1 (CK-1), a type 1 casein kinase. In this study it is shown that CK-1 is potently inhibited by heparin when phosvitin or calcineurin are used as substrates. With casein as a substrate, however, the kinase is insensitive to inhibition by heparin. Using phosvitin as a substrate half-maximal inhibition of CK-1 was observed with 0.14 microgram/ml heparin. Kinetic analyses indicate that at a constant concentration (0.10 mM) of ATP the Km of CK-1 for phosvitin is increased eightfold in the presence of 0.9 microgram/ml heparin; the Vmax is unchanged with or without heparin. At a constant concentration of phosvitin (4 mg/ml) heparin (0.9 microgram/ml) decreased the Vmax for ATP by 57%; the Km is unchanged with or without heparin. The inhibition of CK-1 by heparin can be reversed by KCl (greater than 100 mM). These results indicate that heparin is a potent inhibitor not only of CK-2 but also of CK-1. Hence heparin inhibition can no longer be arbitrarily used as a criterion to discriminate between these kinases.  相似文献   

6.
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   

7.
The mechanism of the heparin-promoted reaction of thrombin with antithrombin III was investigated by using covalent complexes of antithrombin III with either high-affinity heparin (Mr = 15,000) or heparin fragments having an average of 16 and 12 monosaccharide units (Mr = 4,300 and 3,200). The complexes inhibit thrombin in the manner of active site-directed, irreversible inhibitors: (Formula: see text) That is, the inhibition rate of the enzyme is saturable with respect to concentration of complexes. The values determined for Ki = (k-1 + k2)/k1 are 7 nM, 100 nM, and 6 microM when the Mr of the heparin moieties are 15,000, 4,300, 3,200, respectively, whereas k2 (2 S-1) is independent of the heparin chain length. The bimolecular rate constant k2/Ki for intact heparin is 3 X 10(8) M-1 S-1 and the corresponding second order rate constant k1 is 6.7 X 10(8) M-1 S-1, a value greater than that expected for a diffusion-controlled bimolecular reaction. The bimolecular rate constants for the complexes with heparin of Mr = 4,300 and 3,200 are, respectively, 2 X 10(7) M-1 S-1 and 3 X 10(5) M-1 S-1. Active site-blocked thrombin is an antagonist of covalent antithrombin III-heparin complexes: the effect is monophasic and half-maximum at 4 nM of antagonist against the complex with intact heparin, whereas the effect is weaker against complexes with heparin fragments and not monophasic. We conclude that virtually all of the activity of high affinity, high molecular weight heparin depends on binding both thrombin and antithrombin III to heparin, and that the exceptionally high activity of heparin results in part from the capacity of thrombin bound nonspecifically to heparin to diffuse in the dimension of the heparin chain towards bound antithrombin III. Increasing the chain length of heparin results in an increased reaction rate because of a higher probability of interaction between thrombin and heparin in solution.  相似文献   

8.
Heparin inhibits the growth of several cell types in vitro, including bovine pulmonary artery smooth muscle cells (BPASMCs). To understand more about the heparin structure required for endogenous activity, chemically modified derivatives of native heparin and glycol-split heparin, namely, 2-O-desulfonated iduronic/glucuronic acid residues in heparin, and 2-O-desulfonated iduronic residues in glycol-split heparin were prepared. These were assayed for their antiproliferative potency on cultured BPASMCs. All of the 2-O-desulfonated heparin derivatives had significantly decreased less antiproliferative activity on BPASMCs. These results suggest that the 2-O-sulfo group of iduronic acid residues in heparin's major sequence is essential for the antiproliferative properties of heparin. The size of heparin does not affect the growth-inhibitory properties of heparin on BPASMCs at the three dose levels examined.  相似文献   

9.
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   

10.
There is a considerable body of experimental evidence that heparin is superior as an anticoagulant to any prothrombin depressing drugs. Furthermore its lipemia-clearing action affords other benefits which result from the removal of fat from the bloodstream. Important among these beneficial effects is the increased tissue and myocardial oxygen consumption which results from the injection of heparin in atherosclerotic patients.Because of these advantages of heparin over oral anticoagulants, the use of heparin as the sole anticoagulant for three weeks in patients with severe acute myocardial infarction was evaluated as opposed to the customary therapy where heparin is given for several days and then oral anticoagulants are used. The mortality in the dicoumarin treated group was 38 per cent, as compared with 28 per cent in the patients who received only heparin for three weeks.  相似文献   

11.
12.
Heparin cofactor II and antithrombin III are plasma proteins functionally similar in their ability to inhibit thrombin at accelerated rates in the presence of heparin. To further characterize the structural and functional properties of human heparin cofactor II as compared to antithrombin III, we studied the possible significance of arginyl and tryptophanyl residues and the changes in protein structure and activity during guanidinium chloride (GdmCl) denaturation. Both antithrombin and heparin cofactor activities of heparin cofactor II are inactivated by the arginine-specific reagent, 2,3-butanedione. Saturation kinetics are observed during modification and suggest formation of a reversible protease inhibitor-butanedione complex. Quantitation of arginyl residues following butanedione modification shows a loss of about four residues for total inactivation, one of which is essential for antithrombin activity. Arginine-modified heparin cofactor II did not bind to heparin-agarose and implies a role for the other modified arginyl residues during heparin cofactor activity. N-Bromosuccinimide oxidation (20 mol of reagent/mol of protein) of heparin cofactor II results in modification of approximately two tryptophanyl residues with no concomitant loss of heparin cofactor activity. Moreover, there is no enhancement of intrinsic protein fluorescence during heparin binding to the native inhibitor. Circular dichroism measurements show that the structural transition of heparin cofactor II during denaturation is distinctly biphasic, yielding midpoints at 0.6 and 2.6 M GdmCl. Functional protease inhibitory activities are affected to the same extent following denaturation-renaturation at various GdmCl concentrations. The results indicate that arginyl residues are critical for both antithrombin and heparin binding activities. In contrast, tryptophanyl residues are apparently not essential for heparin-dependent interactions. The results also suggest that heparin cofactor II contains two structural domains which unfold at different GdmCl concentrations.  相似文献   

13.
The diversity-oriented chemical modification of heparin is shown to afford charge-reduced heparin derivatives that possess increased selectivity for binding heparin-binding proteins. Variable N-desulfonation of heparin was employed to afford heparin fractions possessing varied levels of free amine. These N-desulfonated heparin fractions were selectively N-acylated with structurally diverse carboxylic acids using a parallel synthesis protocol to generate a library of 133 heparin-derived structures. Screening library members to compare affinity for heparin-binding proteins revealed unique heparin-derived structures possessing increased affinity and selectivity for individual heparin-binding proteins. Moreover, N-sulfo groups in heparin previously shown to be required for heparin to bind specific proteins have been replaced with structurally diverse non-anionic moieties to afford identification of charge-reduced heparin derivatives that bind these proteins with equivalent or increased affinity compared to unmodified heparin. The methods described here outline a process that we feel will be applicable to the systematic chemical modification of natural polyanionic polysaccharides and the preparation of synthetic oligosaccharides to identify charge-reduced high affinity ligands for heparin-binding proteins.  相似文献   

14.
The influence of heparin on the inhibition of factor Xa has been studied under conditions where factor Xa is bound to collagen-thrombin-stimulated platelets to form the prothrombinase complex. Unfractionated heparin was found to cause a concentration-dependent acceleration of the inhibition of the platelet prothrombinase complex up to a maximum rate constant of 4.1 X 10(7) M-1 X min-1 at heparin concentrations of 0.2 microM and above. This is equivalent to a 4800-fold acceleration over the rate constant for the inhibition in the absence of heparin, and is 6.8-fold lower than the rate constant for the inhibition of uncomplexed factor Xa in the presence of saturating concentrations of heparin which was determined as 2.8 X 10(8) M-1 X min-1. The effects of three Mr fractions of heparin were also studied. These were a gel-filtered heparin of Mr 15000, a gel-filtered heparin of Mr 6000 and a heparin oligosaccharide (primarily 8-10 monosaccharide units) prepared by nitrous acid depolymerization, each with high affinity for antithrombin III. These fractions all accelerated the rate of the antithrombin III inhibition of the platelet prothrombinase complex, with maximum rate constants of 6.8 X 10(7), 1.4 X 10(7) and 9.8 X 10(6) M-1 X min-1, respectively. On comparison with the effect of these heparin fractions on the rate of inhibition of uncomplexed factor Xa a progressively increasing disparity between the rate of inhibition of uncomplexed and complexed factor Xa was observed, rising from 1.7-fold with the oligosaccharide to 6.8-fold with the unfractionated heparin. A possible mechanism for this differential activity between uncomplexed and complexed factor Xa with the various heparin fractions is discussed in terms of an involvement of heparin binding to factor Xa.  相似文献   

15.
Studies were conducted to determine the effect of modifying specific functional groups of heparin on its antithrombin III-enhancing activity. The derivatives employed were heparin methyl ester, heparinylglycine and N-desulfated heparin. The carboxyl-modified derivatives increase the rate of inhibition of thrombin by antithrombin III, although not to the same extent as heparin. N-Desulfated heparin is devoid of any activity. Heparin methyl ester is more potent than heparinylglycine in activating antithrombin III, as exhibited by its immediate effect on the thrombin-fibrinogen reaction. However, heparinylglycine is the more effective of the two, in increasing the rate of thrombin deactivation by antithrombin III. The results indicate that although free carboxyl groups of heparin are not crucial for its binding to antithrombin III, they are important for the combination of the latter with thromobin. In contrast, N-sulfates are critical for the interaction of heparin with antithrombin III.  相似文献   

16.
Vitronectin is a 70-kDa protein that is found in both the extracellular matrix as well as serum. Vitronectin is one of the few proteins that regulates both the complement and the coagulation systems. Heparin is known to bind to vitronectin. Review of the literature reveals apparently conflicting outcomes of the interaction of heparin, vitronectin, and the complement system. Previous studies demonstrated that heparin diminishes vitronectin inhibition of complement activity. Numerous studies have also demonstrated that heparin exerts a net inhibitory effect on complement. We used two dimensional affinity resolution electrophoresis (2DARE) to examine this apparent paradox. 2DARE allowed simultaneous determination of binding affinity of heparin for vitronectin as well as the M(r) of the heparin species. In the 2DARE experiment, the interaction of heparin with vitronectin caused retardation of the movement of the heparin through the tube gel in the first dimension. The degree of the retardation of movement was used to calculate the approximate K(d) of that interaction. The heparin from the tube gel was then subjected to a second dimension electrophoresis to determine the M(r) of the heparin. 2DARE analysis of the interaction of heparin with vitronectin clearly demonstrated that a sub-population of heparin chains with M(r) > 8000 bound vitronectin with high affinity whereas most high M(r) chains and all lower M(r) chains showed little to no affinity for vitronectin. Our findings are consistent with the hypothesis that a unique binding domain exists in certain heparin chains for vitronectin.  相似文献   

17.
The 165 amino acid form of vascular endothelial growth factor (VEGF165) is a heparin-binding growth factor with mitogenic activity for vascular endothelial cells. We examined activities of various heparin derivatives toward their interactions with VEGF165 using an enzyme-linked immunosorbent assay and elucidated the structural features in heparin for the interactions. Native heparin interacted with VEGF165, whereas N-desulfated, N-acetylated (N-DS, N-Ac-) heparin, and 6-O-desulfated (6-O-DS-) heparin did not. The 2-O-desulfated (2-O-DS-) heparin retained the ability for the interaction with VEGF165. In contrast, the 2-O-DS-heparin exhibited no ability for the interaction with FGF-2 and HGF. Thus, structural requirements in heparin for the specific interaction with VEGF165 are distinct from those with FGF-2 and HGF which require a high content of 2-O-sulfate groups. In a cell proliferation assay, native heparin and 2-O-DS-heparin exhibited inhibitory abilities for VEGF165-induced proliferation of human umbilical vein endothelial cells (HUVECs) with their high concentrations (more than 64 microg/ml), while only native heparin could enhance the proliferation of the chlorate-treated cells. These results suggested that a high content of 2-O-sulfate groups is not required for the specific interaction with VEGF165alone, although it is essential for the mitogenic activity of the growth factor.  相似文献   

18.
An electrophoretic method for the quantitation and preparation of antithrombin III-high-affinity heparin using agarose beds is described. The method allows the determination of high-affinity heparin fractions in several samples in one single step. The incubation mixture containing heparin and antithrombin III is submitted to agarose gel electrophoresis in 0.06 m barbital buffer, pH 8.6. A sharp separation between free antithrombin III, the complex antithrombin III-heparin, and free heparin occurs under these conditions. Around 30% of heparin molecules present in commerical preparations bind to antithrombin. This bound heparin has an anticoagulant activity of 240 IU. Negligible binding of other sulfated mucopolysaccharides to antithrombin III was observed. The whole procedure takes less than 6 h and can also be used as a semipreparative method for high-affinity heparin.  相似文献   

19.
The serpin heparin cofactor II (HCII) is a glycosaminoglycan-activated inhibitor of thrombin that circulates at a high concentration in the blood. The antithrombotic effect of heparin, however, is due primarily to the specific interaction of a fraction of heparin chains with the related serpin antithrombin (AT). What currently prevents selective therapeutic activation of HCII is the lack of knowledge of the determinants of glycosaminoglycan binding specificity. In this report we investigate the heparin binding properties of HCII and conclude that binding is nonspecific with a minimal heparin length of 13 monosaccharide units required and affinity critically dependent on ionic strength. Rapid kinetics of heparin binding indicate an induced fit mechanism that involves a conformational change in HCII. Thus, HCII binds to heparin in a manner analogous to the interaction of AT with low affinity heparin. A fully allosteric 2000-fold heparin activation of thrombin inhibition by HCII is demonstrated for heparin chains up to 26 monosaccharide units in length. We conclude that the heparin-binding mechanism of HCII is closely analogous to that of AT and that the induced fit mechanism suggests the potential design or discovery of specific HCII agonists.  相似文献   

20.
The molecular mechanism by which heparin modulates the processing of procathepsin L in the extracellular environment is proposed. We show that heparin reduces the stability of the pro form of cathepsin L at pH 5 by binding to a putative heparin binding motif (BBXB) in the pro-domain. Mutations to this motif on procathepsin L reduce heparin binding affinity and heparin-induced destabilization; in contrast, heparin only slightly destabilizes the mature cathepsin L domain. Gel analysis further shows that heparin makes procathepsin L a much better substrate for cathepsin L. Thus, heparin enhances the rate of zymogen activation by destabilization upon binding to the BBXB motif. Determining the mechanism by which procathepsin L is activated in the extracellular matrix is important to the understanding of the role that cathepsin L plays in tumour invasion.  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