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.
The interaction between bovine antithrombin, a plasma proteinase inhibitor, and heparin species of different molecular weights was studied. A commercial heparin preparation was divided by gel chromatography into a number of fractions with average molecular weights ranging from 6000 to 34700. Each of these fractions was further fractionated by affinity chromatography on matrix-bound antithrombin. In the latter procedure, those heparin fractions that had molecular weights lower than about 14000 were separated into three peaks. The material in the first of these was not adsorbed on the column, and the other two peaks corresponded to the low-affinity and high-affinity peaks described previously. In contrast, high-molecular-weight heparin samples gave only the low-affinity and high-affinity fractions. U.v. difference absorption studies showed that the non-adsorbed heparin fraction bound to antithrombin in solution with a binding constant at physiological ionic strength only slightly lower than that of low-affinity heparin. The division between the two fractions thus is arbitrary and only dependent on the conditions selected for the affinity-chromatography experiment. Stoicheiometries and binding constants for the binding of several high-affinity heparin species to antithrombin were determined by fluorescence titrations. High-affinity heparin fractions of equal elution positions in the beginning of the peaks of the affinity chromatographies, but with different molecular weights, showed stoicheiometries that were not experimentally distinguishable from 1:1 and also had no appreciable differences in binding constants. However, the anticoagulant activities, calculated on a molar basis, of these fractions increased markedly with molecular weight, a behaviour that thus cannot be explained by differences in the binding of the fractions to antithrombin. In contrast, high-affinity samples of similar molecular weights, which were eluted at increasing ionic strengths from matrix-linked antithrombin, were found to have an increasing proportion of chains with two binding sites for antithrombin and also to have progressively higher binding constants. These binding properties at least partly explain the increasing anticoagulant activities that were observed for these fractions.  相似文献   

2.
Antithrombin III exists in plasma as major and minor isoforms differing in affinity for heparin. The nature of the binding of each purified isoform to immobilized heparins was investigated. Unfractionated, mixed-affinity heparin bound each isoform with both high affinity and concentration-dependent low affinity. The isoforms were resolved when filtered through low-affinity heparin (heparin repeatedly passed over immobilized antithrombin III) columns. Following chemical modification of a specific tryptophan residue required for heparin binding, each isoform failed to bind to either low-affinity or mixed-affinity heparin-agarose, but elution of the modified higher-affinity isoform was retarded on both gels. Because the modified lower-affinity isoform eluted with the similarly sized bovine serum albumin in these experiments, the difference in isoform affinity for heparin appears to be the result of a unique, secondary heparin-binding site in the higher-affinity isoform that can bind a heparin site with low affinity for antithrombin III. This interpretation was supported by the chromatographic behavior of the isoforms on mixed-affinity agarose during reverse gradient elution. Two other populations of each of the tryptophan-modified isoforms were identified. Since these isoforms bound tightly to mixed-affinity heparin-agarose but eluted at lower salt concentrations than the corresponding unmodified isoforms, both isoforms may contain additional secondary sites that interact weakly with heparin. A general model of heparin-antithrombin III interaction is proposed in which a high-affinity heparin site initially interacts with a primary site on antithrombin III. The subsequent conformational change leads to a cooperative, entropy-driven association between secondary sites on the protein and low-affinity sites on heparin, stabilizing antithrombin III in its activated form.  相似文献   

3.
Oligosaccharides (10-20 monosaccharide units) with high affinity for antithrombin, as well as larger high-affinity heparin fractions (having relative molecular masses between 6,000 and 21,500), all markedly accelerated the inhibition of Factor Xa by antithrombin. Moreover, all high-affinity oligosaccharides and heparins enhanced, to a similar extent, the amount of free proteolytically modified antithrombin cleaved at the reactive bond by Factor Xa. In contrast, a minimum high-affinity heparin size of approximately 18 monosaccharide units was required to significantly accelerate the inactivation of thrombin by antithrombin and to enhance the production of modified antithrombin by this enzyme. All high-affinity fractions studied had similar affinities for antithrombin, as determined by fluorescence titrations. In competition experiments, binary complexes of antithrombin with octadecasaccharide or larger high-affinity heparins, but not with smaller oligosaccharides, displaced inactivated 125I-thrombin from matrix-linked low-affinity heparin. Moreover, similar binary complexes with 3H-labeled octadecasaccharide or larger chains, but not with smaller oligosaccharides, were capable of binding to matrix-linked inactivated thrombin. These results indicate that simultaneous binding of antithrombin and thrombin to high-affinity heparin is a prerequisite to the acceleration of the antithrombin-thrombin reaction and that the minimum heparin sequence capable of binding both proteins comprises approximately 18 monosaccharide units. Similar complex formation apparently is not required for the acceleration of the antithrombin-Factor Xa reaction.  相似文献   

4.
Futamura A  Beechem JM  Gettins PG 《Biochemistry》2001,40(22):6680-6687
Activation of antithrombin by high-affinity heparin as an inhibitor of factor Xa has been ascribed to an allosteric switch between two conformations of the reactive center loop. However, we have previously shown that other, weaker binding, charged polysaccharides can give intermediate degrees of activation [Gettins, P. G. W., et al. (1993) Biochemistry 32, 8385-8389]. To examine whether such intermediate activation results from different reactive center loop conformations or, more simply, from a different equilibrium constant between the same two extreme conformations, we have used NBD covalently bound at the P1 position of an engineered R393C variant of antithrombin as a fluorescent reporter group and measured fluorescence lifetimes of the label in free antithrombin as well as in antithrombin saturated with long-chain high-affinity heparin, high-affinity heparin pentasaccharide, long-chain low-affinity heparin, and dextran sulfate. Steady state emission spectra, anisotropies, and dynamic quenching measurements were also recorded. We found that the large steady state fluorescence enhancements produced by binding of activators resulted from relief of a static quench of fluorescence of NBD in approximately 50% of the labeled antithrombin molecules rather than from any large change in lifetimes, and that similar lifetimes were found for NBD in all activated antithrombin-oligosaccharide complexes. Similar anisotropies and positions of the NBD emission maxima were also found in the absence and presence of activators. In addition, NBD was accessible to quenching agents in both the absence and presence of activators, with an at most 2-fold increase in quenching constants between these two extremes. The simplest interpretation of the partial static quench in the absence of activators, the different degrees of enhancement by different antithrombin activators, and the similar fluorescence properties and quenching behavior of the different states is that there are two distinct types of conformational equilibrium involving three distinct states of antithrombin, which we designate A, A', and B. A and A' represent low-affinity or inactive states of approximately equal energy, both having the hinge residues inserted into beta-sheet A. A is fluorescent, while A' is statically quenched. State B represents the activated loop-expelled conformation in which none of the NBD fluorophores are statically quenched, as a result of the loop, including the P1-NBD, moving away from the body of the antithrombin. Different activators are able to shift the equilibrium to the high-activity (B) state to different extents and hence give different degrees of measured activity, and different degrees of relief of static quench. The similar properties and accessibility of the NBD in the A and B conformations also indicate that the P1 side chain is not buried in the low-activity A conformation, suggesting that an earlier proposal that activation involves exposure of the P1 side chain cannot be the explanation for activation. As an alternative explanation, heparin activation may give access to an exosite on antithrombin for binding to factor Xa and hence be the principal basis for enhancement of the rate of inhibition.  相似文献   

5.
V Arocas  S C Bock  S T Olson  I Bj?rk 《Biochemistry》1999,38(31):10196-10204
Heparin greatly accelerates the reaction between antithrombin and its target proteinases, thrombin and factor Xa, by virtue of a specific pentasaccharide sequence of heparin binding to antithrombin. The binding occurs in two steps, an initial weak interaction inducing a conformational change of antithrombin that increases the affinity for heparin and activates the inhibitor. Arg46 and Arg47 of antithrombin have been implicated in heparin binding by studies of natural and recombinant variants and by the crystal structure of a pentasaccharide-antithrombin complex. We have mutated these two residues to Ala or His to determine their role in the heparin-binding mechanism. The dissociation constants for the binding of both full-length heparin and pentasaccharide to the R46A and R47H variants were increased 3-4-fold and 20-30-fold, respectively, at pH 7.4. Arg46 thus contributes only little to the binding, whereas Arg47 is of appreciable importance. The ionic strength dependence of the dissociation constant for pentasaccharide binding to the R47H variant showed that the decrease in affinity was due to the loss of both one charge interaction and nonionic interactions. Rapid-kinetics studies further revealed that the affinity loss was caused by both a somewhat lower forward rate constant and a greater reverse rate constant of the conformational change step, while the affinity of the initial binding step was unaffected. Arg47 is thus not involved in the initial weak binding of heparin to antithrombin but is important for the heparin-induced conformational change. These results are in agreement with a previously proposed model, in which an initial low-affinity binding of the nonreducing-end trisaccharide of the heparin pentasaccharide induces the antithrombin conformational change. This change positions Arg47 and other residues for optimal interaction with the reducing-end disaccharide, thereby locking the inhibitor in the activated state.  相似文献   

6.
Alignment of the heparin-activated serpins indicates the presence of two binding sites for heparin: a small high-affinity site on the D-helix corresponding in size to the minimal pentasaccharide heparin, and a longer contiguous low-affinity site extending to the reactive center pole of the molecule. Studies of the complexing of antithrombin and its variants with heparin fractions and with reactive center loop peptides including intermolecular loop-sheet polymers all support a 3-fold mechanism for the heparin activation of antithrombin. Binding to the pentasaccharide site induces a conformational change as measured by circular dichroism. Accompanying this, the reactive center becomes more accessible to proteolytic cleavage and there is a 100-fold increase in the kass for factor Xa but only a 10-fold increase for thrombin, to 6.4 x 10(4) M-1 s-1. To obtain a 100-fold increase in the kass for thrombin requires in addition a 4:1 molar ratio of disaccharide to neutralize the charge on the extended low-affinity site. Full activation requires longer heparin chains in order to stabilize the ternary complex between antithrombin and thrombin. Thus, addition of low-affinity but high molecular weight heparin in conjunction with pentasaccharide gives an overall kass of 2.7 x 10(6) M-1 s-1, close to that of maximal heparin activation.  相似文献   

7.
Equilibrium gel permeation chromatography was employed to determine the ability of heparin to form complexes with thrombin and antithrombin III. In the eluate from a Sephacryl S-200 column, heparin caused a peak and then a trough in the fluorescence of 48 nM antithrombin III or 63 nM thrombin. The peak-heights with known amounts of heparin were used for standard curves to determine the extent of complex formation by test heparin preparations. Only heparin species with high-affinity for antithrombin III specifically formed a complex with antithrombin III under the conditions used. The ability to form a complex of heparin preparations with different anticoagulant activities for thrombin and antithrombin III could be determined satisfactorily. The heparin species with different affinities for antithrombin III did not coincide those with different affinities for thrombin. Of 4 preparations with one low-affinity and three high-affinity subfractions of heparin for antithrombin III, the species with the lowest affinity for antithrombin III had the highest affinity for thrombin. All of these observations showed that the method could be used to determine the ability to form a complex of test heparin preparations.  相似文献   

8.
We have previously shown that heparin is a potent inhibitor of a mammalian DNA topoisomerase I. We have now investigated the mechanism of its inhibition. This was carried out first by scrutinizing the structural features of heparin molecules responsible for the inhibition. Commercial heparin preparation was fractionated by antithrombin III-Sepharose into non-adsorbed, low-affinity and high-affinity fractions, of which only the high-affinity fraction of heparin is known to contain a specific oligosaccharide sequence responsible for the binding to antithrombin III. These fractions all exhibited essentially similar inhibitory activities. Furthermore, when chemically sulphated to an extent comparable with or higher than heparin, otherwise inactive glycosaminoglycans such as heparan sulphate, chondroitin 4-sulphate, dermatan sulphate and neutral polysaccharides such as dextran and amylose were converted into potent inhibitors. Sulphated dermatan sulphate, one of the model compounds, was further shown to bind competitively to the same sites on the enzyme as heparin. These observations strongly suggested that topoisomerase inhibition by heparin is attributable primarily, if not entirely, to the highly sulphated polyanionic nature of the molecules. In a second series of experiments we examined whether heparin inhibits only one or both of the topoisomerase reactions, i.e. nicking and re-joining. It was demonstrated that both reactions were inhibited by heparin, but the nicking reaction was more severely affected than was the re-joining reaction.  相似文献   

9.
From structural analysis on genetically abnormal and chemically modified human antithrombin III [Koide, T., Odani, S., Takahashi, K., Ono, T. and Sakuragawa, N. (1984) Proc. Natl Acad. Sci. USA 81, 289-293; Chang, J.-Y. and Tran, T. H., (1986) J. Biol. Chem. 261, 1174-1176; Blackburn, M. N., Smith, R. L., Carson, J. and Sibley, C. C. (1984) J. Biol. Chem. 259, 939-941], the heparin-binding site of antithrombin III has been suggested to be in the region of Pro-41, Arg-47 and Trp-49. In this study the heparin-binding site was probed by preferential cleavage of V8 protease on heparin-treated and non-treated native antithrombin III. The study has been based on the presumption that the heparin-binding site of antithrombin III is situated at exposed surface domain and may be preferentially attacked during limited proteolytic digestion. Partially digested antithrombin III samples were monitored by quantitative amino-terminal analysis and amino acid sequencing to identify the preferential cleavage sites. 1-h-digested antithrombin III was separated on HPLC and peptide fragments were isolated and characterized both qualitatively and quantitatively. The results reveal that Glu-Gly (residues 34-35), Glu-Ala (residues 42-43) and Glu-Leu (residues 50-51) are three preferential cleavage sites for V8 protease and their cleavage, especially the Glu-Ala and the Glu-Leu sites, was drastically inhibited when antithrombin III was preincubated with heparin. Both high-affinity and low-affinity antithrombin-III-binding heparins were shown to inhibit the V8 protease digestion of native antithrombin III, but the high-affinity sample exhibited a higher inhibition activity than the low-affinity heparin. These findings (a) imply that the segment containing residues 34-51 is among the most exposed region of native antithrombin III and (b) support the previous conclusions that this region may play a pivotal role in the heparin binding.  相似文献   

10.
We evaluated the role of nonspecific electrostatic binding in the interaction of antithrombin (AT) with heparin (Hp), a paradigmatic protein-glycosaminoglycan (GAG) system. To do so, we obtained the ionic-strength dependence of the binding constant, since a common feature in protein-polyelectrolyte systems is a maximum in affinity in the ionic strength range 10 mM 相似文献   

11.
Arocas V  Turk B  Bock SC  Olson ST  Björk I 《Biochemistry》2000,39(29):8512-8518
The interaction of a well-defined pentasaccharide sequence of heparin with a specific binding site on antithrombin activates the inhibitor through a conformational change. This change increases the rate of antithrombin inhibition of factor Xa, whereas acceleration of thrombin inhibition requires binding of both inhibitor and proteinase to the same heparin chain. An extended heparin binding site of antithrombin outside the specific pentasaccharide site has been proposed to account for the higher affinity of the inhibitor for full-length heparin chains by interacting with saccharides adjacent to the pentasaccharide sequence. To resolve conflicting evidence regarding the roles of Lys136 and Lys139 in this extended site, we have mutated the two residues to Ala or Gln. Mutation of Lys136 decreased the antithrombin affinity for full-length heparin by at least 5-fold but minimally altered the affinity for the pentasaccharide. As a result, the full-length heparin and pentasaccharide affinities were comparable. The reduced affinity for full-length heparin was associated with the loss of one ionic interaction and was caused by both a lower overall association rate constant and a higher overall dissociation rate constant. In contrast, mutation of Lys139 affected neither full-length heparin nor pentasaccharide affinity. The rate constants for inhibition of thrombin and factor Xa by the complexes between antithrombin and full-length heparin or pentasaccharide were unaffected by both mutations, indicating that neither Lys136 nor Lys139 is involved in heparin activation of the inhibitor. Together, these results show that Lys136 forms part of the extended heparin binding site of antithrombin that participates in the binding of full-length heparin chains, whereas Lys139 is located outside this site.  相似文献   

12.
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   

13.
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   

14.
The limiting concentrations, i.e., those concentrations of sodium chloride required to completely disrupt the complexes of heparin with antithrombin III, protamine and poly(l-lysine), were determined using fluorescence techniques, in order to compare the binding strengths of these complexes. From the limiting salt concentration values, poly(l-lysine)_always exhibited stronger binding to heparin of a particular anticoagulant potentcy (degree of sulphation) than did protamine. The binding strengths of both complexes decreased as the degree of sulphation of the heparin participating in the complex was reduced. In contrast, the limiting salt concentration values for complexes formed between antithrombin III and heparin did not change with either the degree of sulphation or the biological potency of the heparin samples. A low-potency heparin simply contained a smaller amount of molecules which possessed the intact antithrombin III binding site (thus being fully ‘anticoagulant active’) than a high-potency sample. Low-affinity heparin did not contain these binding sites and thus showed a low affinity for antithrombin III. High-potency heparin, being highly sulphated, possessed a higher affinity for protamine and poly(l-lysine) than for antithrombin III. However, after partial N-desulphation of heparin, the subsequent heparin-protamine complex was more weakly bound than a significant proportion of the corresponding heparin-antithrombin III complexes. These in vitro findinds may have particular relevance in relation to the clinical condition termed ‘hheparin rebound’.  相似文献   

15.
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   

16.
X J Sun  J Y Chang 《Biochemistry》1990,29(38):8957-8962
Arginyl residues of human antithrombin III have been implicated to involve in the heparin binding site [Jorgensen, A. M., Borders, C. L., & Fish, W. W. (1985) Biochem, J. 231, 59-63]. We have performed chemical modification of antithrombin with (p-hydroxyphenyl)glyoxal (HPG) in order to determine the locations of these arginine residues. Antithrombin was modified with 12 mM HPG in the absence and presence of heparin (2-fold by weight to antithrombin). In the absence of heparin, about 3-4 mol of arginines/mol of antithrombin were modified within 60 min, and the modification led to the loss of 95% of the inhibitor's heparin cofactor activity as well as heparin-induced fluorescence enhancement and 50% of its progressive inhibitory activity. In the presence of heparin, the extent of modification was diminished by 30% and modified antithrombin retained approximately 70% of its heparin cofactor activity. Peptide mapping and subsequent sequence analysis revealed that selective HPG modification occurred at Arg129 and Arg145 and that their modifications were protected upon binding of heparin to antithrombin. We conclude that Arg129 and Arg145 are situated within the heparin binding site of human antithrombin III.  相似文献   

17.
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   

18.
Mechanism of the anticoagulant action of heparin   总被引:20,自引:0,他引:20  
Summary The anticoagulant effect of heparin, a sulfated glycosaminoglycan produced by mast cells, requires the participation of the plasma protease inhibitor antithrombin, also called heparin cofactor. Antithrombin inhibits coagulation proteases by forming equimolar, stable complexes with the enzymes. The formation of these complexes involves the attack by the enzyme of a specific Arg-Ser bond in the carboxy-terminal region of the inhibitor. The complexes so formed are not dissociated by denaturing solvents, which indicates that a covalent bond may contribute to their stability. This bond may be an acyl bond between the active-site serine of the enzyme and the arginine of the cleaved reactive bond of the inhibitor. However, the native complexes dissociate slowly at near-neutral pH into free enzyme and a modified inhibitor, cleaved at the reactive bond. So, antithrombin apparently functions as a pseudo-substrate that traps the enzyme in a kinetically stable complex.The reactions between antithrombin and coagulation proteases are slow in the absence of heparin. However, optimal amounts of heparin accelerate these reactions up to 2 000-fold, thereby efficiently preventing the formation of fibrin in blood. The accelerating effect, and thus the anticoagulant activity, is shown by only about one-third of the molecules in all heparin preparations, while the remaining molecules are almost inactive. The highly active molecules bind tightly to antithrombin, i.e. with a binding constant of slightly below 108 M–1 at physiological ionic strength, while the relatively inactive molecules bind about a thousand-fold more weakly. The binding of the high-affinity heparin to antithrombin is accompanied by a conformational change in the inhibitor that is detectable by spectroscopic and kinetic methods. This conformational change follows an initial, weak binding of heparin to antithrombin and causes the tight interaction between polysaccharide and inhibitor that is prerequisite to heparin anticoagulant activity. It has also been postulated that the conformational change leads to a more favourable exposure of the reactive site of antithrombin, thereby allowing the rapid interaction with the proteases.Heparin also binds to the coagulation proteases. Recent studies indicate that this binding is weaker and less specific that the binding to antithrombin. Nevertheless, for some enzymes, thrombin, Factor IXa and Factor XIa, an interaction between heparin and the protease, in addition to that between the polysaccharide and antithrombin; apparently is involved in the accelerated inhibition of the enzymes. The effect of this interaction may be to approximate enzyme with inhibitor in an appropriate manner. However, the bulk of the evidence available indicates that binding of heparin to the protease alone cannot be responsible for the accelerating effect of the polysaccharide on the antithrombin-protease reaction.Heparin acts as a catalyst in the antithrombin-protease reaction, i.e. it accelerates the reaction in non-stoichiometric amounts and is not consumed during the reaction. This ability can be explained by heparin being released from the antithrombin-protease complex for renewed binding to antithrombin, once the complex has been formed. Such a decresed affinity of heparin for the antithrombin complex, compared to the affinity for antithrombin alone, has been demonstrated.The structure of the antithrombin-binding region in heparin has been investigated following the isolation of oligosaccharides with high affinity for antithrombin. The smallest such oligosaccharide, an octasaccharide, obtained after partial random depolymerization of heparin with nitrous acid, was found to contain a unique glucosamine-3-O-sulfate group, which could not be detected in other portions of the high affinity heparin molecule and which was absent in heparin with low affinity for antithrombin. The actual antithrombin-binding region within this octasaccharide molecule has been identified as a pentasaccharide sequence with he predominant structure: N-acetyl-D-glucosamine(6-O-SO3)D-glucoronic acidD-glucosamine(N-SO3;3,6-di-O-SO3)L-iduronic acid(2-O-SO3)D-glucosamine(N-SO3;6-O-SO3). In addition to the 3-O-sulfate group, both N-sulfate groups as well as the 6-O-sulfate group of the N-acetylated glucosamine unit appear to be essential for the interaction with antithrombin. The remarkably constant structure of this sequence, as compared to other regions of the heparin molecule, suggests a strictly regulated mechanism of biosynthesis.The ability of heparin to potentiate the inhibition of blood coagulation by antithrombin generally decreases with decreasing molecular weight of the polysaccharide. However, individual coagulation enzymes differ markedly with regard to this molecular-weight dependence. Oligosaccharides in the extreme low-molecular weight range, i.e. octa- to dodecasaccharides, with high affinity for antithrombin have high anti-Factor Xa-activity but are virtually unable to potentiate the inhibition of thrombin. Furthermore, such oligosaccharides are ineffective in preventing experimentally induced venous thrombosis in rabbits. Slightly larger oligosaccharides, containing 16 to 18 monosaccharide residues, show significant anti-thrombin as well as antithrombotic activities, yet have little effect on overall blood coagulation. These findings indicate that the affinity of a heparin fragment for antithrombin is not in itself a measure of the ability to prevent venous thrombo-genesis, and that the anti-Factor Xa activity of heparin is only a partial expression of its therapeutic potential as an antithrombotic agent.The biological role of the interaction between heparin and antithrombin is unclear. In addition to a possible function in the regulation of hemostasis, endogenous heparin may serve as a regulator of extravascular serine proteinases. Mouse peritoneal macrophages have been found to synthesize all the enzymes that constitute the extrinsic pathway of coagulation. Moreover, tissue thromboplastin is produced by these cells in response to a functional interaction with activated T-lymphocytes. The inhibition of this extravascular coagulation system by heparin, released from mast cells, may be potentially important in modulating inflammatory reactions.  相似文献   

19.
It is proposed that the anti-coagulant activity of heparin is related to the probability of finding, in a random distribution of different disaccharides, a dodecasaccharide with the sequence required for binding to antithrombin. It is shown that this probability is a function of the degree of polymerization of heparin. The hypothesis has been been tested with a series of narrow-molecular-weight-range fractions ranging from 5,600 to 36,000. The fractions having mol.wts. below 18,000 (comprising 85% of the original preparation) followed the predicted probability relationship as expressed by the proportion of molecules capable of binding to antithrombin. The probability that any randomly chosen dodecasaccharide sequence in heparin should bind to antithrombin was calculated to 0.022. The fraction with mol.wt. 36,000 contained proteoglycan link-region fragments, which may explain the deviation of the high-molecular-weight fractions from the hypothetical relationship. The relationship between anti-coagulant activity and molecular weight cannot be explained solely on the basis of availability of binding sites for antithrombin. The activity of high-affinity heparin (i.e. molecules containing high-affinity binding sites for antithrombin), determined either by a whole-blood clotting procedure or by thrombin inactivation in the presence of antithrombin, thus remained dependent on molecular weight. Possible explanations of this finding are discussed. One explanation could be a requirement for binding of thrombin to the heparin chain adjacent to antithrombin.  相似文献   

20.
Specific binding of the anticoagulants heparin and antithrombin III to the blood clotting cascade factor human thrombin was recorded as a function of time with a Love-wave biosensor array consisting of five sensor elements. Two of the sensor elements were used as references. Three sensor elements were coated with RNA or DNA aptamers for specific binding of human thrombin. The affinity between the aptamers and thrombin, measured using the biosensor, was within the same range as the value of K(D) measured by filter binding experiments. Consecutive binding of the thrombin inhibitors heparin, antithrombin III or the heparin-antithrombin III complex to the immobilized thrombin molecules, and binding of a ternary complex of heparin, anithrombin III, and thrombin to aptamers was evaluated. The experiments showed attenuation of binding to thrombin due to heparin-antithrombin III complex formation. Binding of heparin activated the formation of the inhibitory complex of antithrombin III with thrombin about 2.7-fold. Binding of the DNA aptamer to exosite II appeared to inhibit heparin binding to exosite I.  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