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.
1. We have used horseradish peroxidase-conjugated protein A- and 125I-protein A to develop immunohistochemical and radioimmunohistochemical methods for the localization of antigens in brain and other tissues of the rat. 2. We visualized methionine-enkephalin fibers in the rat brain by incubating tissue sections with a specific polyclonal antibody and peroxidase-conjugated protein A. The method is simple, fast, and less expensive and more sensitive than classical immunohistochemical techniques and the principle could be used to visualize many other tissue antigens. 3. Incubation of tissue samples with specific polyclonal antibodies and 125I-protein A, followed by autoradiography, allows the permanent recording of the radioimmunohistochemical localization of brain methionine-enkephalin, tyrosine hydroxylase, and angiotensin-converting enzyme and of pituitary vasopressin and could be applied to the localization of many other tissue antigens. 4. A new quantitative radioimmunohistochemical technique for methionine-enkephalin allows the determination of the endogenous peptide content in discrete brain nuclei from 16-microns-thick sections. The method is based on the quantitative determination of the amount of 125I-protein A bound to specific tissue areas after incubation with a specific polyclonal antibody, followed by autoradiography and computerized microdensitometry. To quantify the endogenous peptide content, the values obtained are interpolated into a methionine-enkephalin internal standard curve. This standard curve was constructed by measuring endogenous concentrations of methionine-enkephalin by radioimmunoassay in specific brain regions and correlating these values with quantitative autoradiographic determinations in homologous areas of adjacent sections. Similar methods can be developed for other tissue antigens. 5. These new methods allow for the localization and quantification of tissue antigens in very discrete areas of the brain and other tissues and have a wide application in neurobiology and pathology.  相似文献   

2.
Hepatic and extrahepatic uptake of intravenously injected lipoprotein lipase   总被引:11,自引:0,他引:11  
Rats were injected intravenously with 125I-labeled bovine lipoprotein lipase. The lipase disappeared within minutes from the blood due to uptake both in the liver (about 50% of the injected dose) and in extrahepatic tissues. Lipase enzyme activity disappeared in parallel to the 125I radioactivity. Thus, there was no inactivation of lipase in the circulating blood. Similar results were obtained when lipoprotein lipase purified from guinea pigs was injected into guinea pigs. Using supradiphragmatic rats we could show that the extrahepatic uptake was saturable and that the amounts of lipase that could be bound far exceeded the amounts of endogenous lipase expected to be present on the endothelium. When the lipase was denatured before injection, its removal in supradiaphragmatic rats became slower, and in intact rats the fraction of the uptake that occurred in extrahepatic tissues was much decreased. It is concluded that recognition by the extrahepatic receptors depends on the native conformation of the lipase. The extrahepatic uptake was strongly impeded by injection of heparin prior to injection of the lipase, and the uptake could to a large extent be reversed by injection of heparin after the lipase. Even after 1 h lipase that had been taken up by extrahepatic tissues reappeared immediately in the blood on injection of heparin. This was true both for enzyme activity and for enzyme radioactivity. Thus, internalization-inactivation-degradation occur only slowly in extrahepatic tissues. It is possible that the extrahepatic binding occurs to the enzyme's physiological receptors. The hepatic uptake was not dependent on the native conformation of the lipase, was less sensitive to heparin, could not be reversed by heparin and was not saturable. The enzyme was not rapidly inactivated after uptake; its activity could be detected in liver homogenates even after 1 h. Degradation to acid-soluble products in the liver was relatively slow; the t1/2 for native lipase was about 1 h. In comparison, in parallel experiments asialofetuin was degraded with a t1/2 of about 15 min.  相似文献   

3.
The effects of autolysis on the structure of chicken calpain II.   总被引:2,自引:2,他引:0       下载免费PDF全文
Heparin catalyses the inhibition of two key enzymes of blood coagulation, namely Factor Xa and thrombin, by enhancing the antiproteinase activities of plasma antithrombin III and heparin cofactor II. In addition, heparin can directly inhibit the activation of Factor X and prothrombin. The contributions of each of these effects to the anticoagulant activity of heparin have not been delineated. We therefore performed experiments to assess how each of these effects of heparin contributes to its anticoagulant activity by comparing the effects of heparin, pentosan polysulphate and D-Phe-Pro-Arg-CH2Cl on the intrinsic pathway of coagulation. Unlike heparin, pentosan polysulphate catalyses only the inhibition of thrombin by plasma. D-Phe-Pro-Arg-CH2Cl is rapid enough an inhibitor of thrombin so that when added to plasma no complexes of thrombin with its inhibitors are formed, whether or not the plasma also contains heparin. Heparin (0.66 microgram/ml) and pentosan polysulphate (6.6 micrograms/ml) completely inhibited the intrinsic-pathway activation of 125I-prothrombin to 125I-prothrombin fragment 1 + 2 and 125I-thrombin. On the addition of thrombin, a good Factor V activator, to the plasma before each sulphated polysaccharide, the inhibition of prothrombin activation was demonstrable only in the presence of higher concentrations of the sulphated polysaccharide. D-Phe-Pro-Arg-CH2Cl also completely inhibited the intrinsic-pathway activation of prothrombin in normal plasma. The inhibitory effect of D-Phe-Pro-Arg-CH2Cl was reversed if thrombin was added to the plasma before D-Phe-Pro-Arg-CH2Cl. The inhibition of the activation of prothrombin by the three agents was also abolished with longer times with re-added Ca2+. Reversal of the inhibitory effects of heparin and pentosan polysulphate was associated with the accelerated formation of 125I-thrombin-antithrombin III and 125I-thrombin-heparin cofactor complexes respectively. These results suggest that the anticoagulant effects of heparin and pentosan polysulphate are mediated primarily by their ability to inhibit the thrombin-dependent activation of Factor V, thereby inhibiting the formation of prothrombinase complex, the physiological activator of prothrombin.  相似文献   

4.
A rat platelet factor has a high antiheparin activity. It also decreases nonenzymatic fibrinolytic activity of normal rat plasma and antithrombin III-heparin complex. The platelet factor 4 formed inactive complexes with heparin in molar ratios of 1:1 and 2:1. Intravenous injection of the platelet factor 4 before injection of albino rats with tissue thromboplastin prevented the reaction of anticoagulation system inactivated the synthesis of endogenous thrombin. This effect is accompanied by high hypercoagulation and depression of nonenzymatic fibrinolysis in blood.  相似文献   

5.
Antithrombotic and clotting factors have long been targets for drug discovery, necessitating the development of blood assays to determine the efficacy of lead compounds prior to animal testing. We have developed a reconstituted blood clot lysis assay which eliminates the need for on-site donors. The assay uses whole blood stored at 4 degrees C obtained from a local blood bank, diluted 1:10 in phosphate buffer. This blood was supplemented with 125I-labeled fibrinogen and the release of radioactive fibrinopeptides from formed clots was measured. The whole blood used in this assay, which had been stored at 4 degrees C for several days, no longer formed solid or retracting clots. Thus, platelets 5-7 days ex vivo (165 x 10(6) platelets) were added to the whole blood in the presence of thrombin (0.80 IU/ml) to form clots. Solid clots formed within 2 min of thrombin addition and began retracting shortly thereafter. In the absence of any thrombolytic agent, clots fully retracted within 2.5 h and remained stable. Thrombin-stimulated clot formation was completely inhibited by heparin. Clots could be lysed in a dose-dependent fashion in the presence of tissue-type plasminogen activator. Clot lysis could be completely inhibited in a dose-dependent fashion with plasminogen activator inhibitor type 1. To demonstrate the utility of this assay as a screen for thrombolytic agents, a 14-amino-acid PAI-1-inhibitory peptide relieved the PAI-1 effect on tPA in a dose-dependent fashion. These data describe an assay for the screening of potential pro-fibrinolytic agents that target PAI-1 inhibition in a human plasma-based system that is versatile, cost-effective, and physiologically relevant and does not rely on the availability of on-site blood donors.  相似文献   

6.
We have identified a tissue-kallikrein-binding protein in human serum and in the serum-free culture media from human lung fibroblasts (WI-38) and rodent neuroblastoma X glioma hybrid cells (NG108-15). Purified and 125I-labelled tissue kallikrein and human serum form an approximately 92,000-Mr SDS-stable complex. The relative quantity of this complex-formation is measured by densitometric scanning of autoradiograms. Complex-formation between tissue kallikrein and the serum binding protein was time-dependent and detectable after 5 min incubation at 37 degrees C, with half-maximal binding at 28 min. Binding of 125I-kallikrein to kallikrein-binding protein is temperature-dependent and can be inhibited by heparin or excess unlabelled tissue kallikrein but not by plasma kallikrein, collagenase, thrombin, urokinase, alpha 1-antitrypsin or kininogens. The kallikrein-binding protein is acid- and heat-labile, as pretreatment of sera at pH 3.0 or at 60 degrees C for 30 min diminishes complex-formation. However, the formed complexes are stable to acid or 1 M-hydroxylamine treatment and can only be partially dissociated with 10 mM-NaOH. When kallikrein was inhibited by the active-site-labelling reagents phenylmethanesulphonyl fluoride or D-Phe-D-Phe-L-Arg-CH2Cl no complex-formation was observed. An endogenous approximately 92,000-Mr kallikrein-kallikrein-binding protein complex was isolated from normal human serum by using a human tissue kallikrein-agarose affinity column. These complexes were recognized by anti-(human tissue kallikrein) antibodies, but not by anti-alpha 1-antitrypsin serum, in Western-blot analyses. The results show that the kallikrein-binding protein is distinct from alpha 1-antitrypsin and is not identifiable with any of the well-characterized plasma proteinase inhibitors such as alpha 2-macroglobulin, inter-alpha-trypsin inhibitor, C1-inactivator or antithrombin III. The functional role of this kallikrein-binding protein and its impact on kallikrein activity or metabolism in vivo remain to be investigated.  相似文献   

7.
Individuals with blood group O have a higher bleeding risk than non-O blood groups. This could be explained by the lower levels of FVIII and von Willebrand Factor (VWF) levels in O individuals. We investigated the relationship between blood groups, thrombin generation (TG), prothrombin activation and thrombin inactivation. Plasma levels of VWF, FVIII, antithrombin, fibrinogen, prothrombin and α2Macroglobulin (α2M) levels were determined. TG was measured in platelet rich (PRP) and platelet poor plasma (PPP) of 217 healthy donors and prothrombin conversion and thrombin inactivation were calculated. VWF and FVIII levels were lower (75% and 78%) and α2M levels were higher (125%) in the O group. TG is 10% lower in the O group in PPP and PRP. Less prothrombin was converted in the O group (86%) and the thrombin decay capacity was lower as well. In the O group, α2M plays a significantly larger role in the inhibition of thrombin (126%). In conclusion, TG is lower in the O group due to lower prothrombin conversion, and a larger contribution of α2M to thrombin inactivation. The former is unrelated to platelet function because it is similar in PRP and PPP, but can be explained by the lower levels of FVIII.  相似文献   

8.
The binding of 125I-transforming growth factors-beta 1 and beta 2 (TGF-beta 1 and TGF-beta 2) to alpha 2-macroglobulin (alpha 2M) was studied before and after reaction with plasmin, thrombin, trypsin, or methylamine. Complex formation between TGF-beta and native or reacted forms of alpha 2M was demonstrated by non-denaturing polyacrylamide gel electrophoresis and autoradiography. Reaction of native alpha 2M with plasmin or methylamine markedly increased the binding of 125I-TGF-beta 1 and 125I-TGF-beta 2 to alpha 2M. The alpha 2M-plasmin/TGF-beta complexes were minimally dissociated by heparin. Reaction of alpha 2M with thrombin or trypsin reduced the binding of 125I-TGF-beta 1 and 125I-TGF-beta 2; the resulting complexes were readily dissociated by heparin. Complexes between TGF-beta 2 and native or reacted forms of alpha 2M were less dissociable by heparin than the equivalent complexes with TGF-beta 1. These studies demonstrate that the TGF-beta-binding activity of alpha 2M is significantly affected by plasmin, thrombin, trypsin and methylamine. Observations that alpha 2M-plasmin preferentially binds TGFs-beta suggest a mechanism by which alpha 2M may regulate availability of TGFs-beta to target cells in vivo.  相似文献   

9.
The inhibition of thrombin by antithrombin III (AT III) and heparin has been studied in pure systems to determine the kinetics of inhibition during human prothrombin activation. The present study shows that prothrombinase-catalyzed prothrombin activation resulted in the generation of thrombin and meizothrombin(des F1). In the absence of heparin the second-order rate constants of the inactivation of both thrombin and meizothrombin(des F1) formed in the reaction mixture appeared to be identical, k = 3.7 X 10(5) M-1 min-1. The rate constant of inhibition of purified thrombin was 6.5 X 10(5) M-1 min-1. In the presence of heparin the decay of the amidolytic activity was biexponential and could be modeled by a four-parameter equation to determine the pseudo first-order rate constants of inhibition as well as the composition of the reaction with respect to the levels of thrombin and meizothrombin(des F1). The ratio of thrombin over meizothrombin(des F1) varied with the initial prothrombin concentration. Heparin catalyzed the AT III inhibition of thrombin but not meizothrombin(des F1) formed during the prothrombin activation. Thrombin, generated by (Xa-Va-phospholipid-Ca2+) was inhibited by AT III/heparin more slowly than purified thrombin, and the saturation kinetics of the inhibition with respect to AT III differed from those found with purified thrombin.  相似文献   

10.
Heparin conjugated polylactide as a blood compatible material   总被引:7,自引:0,他引:7  
Jee KS  Park HD  Park KD  Kim YH  Shin JW 《Biomacromolecules》2004,5(5):1877-1881
A heparin-conjugated biodegradable polymer (PLA-heparin) by the direct coupling of heparin to polylactide (PLA) was synthesized and characterized. The surface exposed heparin content associated PLA-heparin was measured to be 0.067 microg/cm2. PLA-heparin coated surface has shown higher hydrophilicity rather than control PLA surface. The clotting time of PLA-heparin conjugate measured by activated partial thromboplastin time (APTT) was significantly prolonged as compared to PLA. The bioactivity of bound heparin measured by APTT corresponds to 17.4% of free heparin. It has been also demonstrated that the conjugation of heparin suppresses the protein adsorption as well as the platelet adhesion. These results indicate that the unique property of bound heparin has an inhibiting influence on the coagulation, plasma protein adsorption, and subsequent platelet adhesion systems. This novel PLA-heparin conjugate could be applied as blood/tissue compatible biodegradable materials for implantable medical devices and tissue engineering.  相似文献   

11.
Heparin cofactor II (Mr = 65,600) was purified 1800-fold from human plasma to further characterize the structural and functional properties of the protein as they compare to antithrombin III (Mr = 56,600). Heparin cofactor II and antithrombin III are functionally similar in that both proteins have been shown to inhibit thrombin at accelerated rates in the presence of heparin. There was little evidence for structural homology between heparin cofactor II and antithrombin III when high performance liquid chromatography-tryptic peptide maps and NH2-terminal sequences were compared. A partially degraded form of heparin cofactor II was also obtained in which a significant portion (Mr = 8,000) of the NH2 terminus was missing. The rates of thrombin inhibition (+/- heparin) by native and partially degraded-heparin cofactor II were not significantly different, suggesting that the NH2-terminal region of the protein is not essential either for heparin binding or for thrombin inhibition. A significant degree of similarity was found in the COOH-terminal regions of the proteins when the primary structures of the reactive site peptides, i.e. the peptides which are COOH-terminal to the reactive site peptide bonds cleaved by thrombin, were compared. Of the 36 residues identified, 19 residues in the reactive site peptide sequence of heparin cofactor II could be aligned with residues in the reactive site peptide from antithrombin III. While the similarities in primary structure suggest that heparin cofactor II may be an additional member of the superfamily of proteins consisting of antithrombin III, alpha 1-antitrypsin, alpha 1-antichymotrypsin and ovalbumin, the differences in structure could account for differences in protease specificity and reactivity toward thrombin. In particular, a disulfide bond which links the COOH-terminal (reactive site) region of antithrombin III to the remainder of the molecule and is important for the heparin-induced conformational change in the protein and high affinity binding of heparin does not appear to exist in heparin cofactor II. This observation provides an initial indication that while the reported kinetic mechanisms of action of heparin in accelerating the heparin cofactor II/thrombin and antithrombin III/thrombin reactions are similar, the mechanisms and effects of heparin binding to the two inhibitors may be different.  相似文献   

12.
The activation of protein C by thrombin is thought to occur at the endothelial cell surface in the presence of an essential membrane glycoprotein cofactor, thrombomodulin. In the present study it is demonstrated that, in the presence of hirudin, the most potent known inhibitor of thrombin, human protein C can be activated by human factor Xa (20 nM), but by a thrombomodulin-independent mechanism requiring only the presence of Ca2+ and phospholipid vesicles bearing a high proportion of negative charges (30-75% phosphatidylserine, depending on the conditions). At an optimal concentration of phosphatidylserine/phosphatidylcholine (1:1, w/w) of 75 microM, the apparent Km was 1 microM with a kcat. of 1 min-1. At 25 microM-phospholipid the Km was unchanged and the kcat. was 0.67 min-1. At either lipid concentration, increasing the density of negative charges by the adjunction of sulphated polysaccharides, like pentosan polysulphate or standard heparin at optimal concentrations of 2-5 micrograms/ml and 5-10 micrograms/ml respectively, resulted in a 4-fold increase of the kcat. without affecting the Km. Sulphated polysaccharides alone were poor promoters of protein C activation by factor Xa. In any case the presence of Ca2+ was essential, the dependence being sigmoidal with Hill coefficients ranging from 1.4 to 2.0. No significant activation of 4-carboxyglutamic acid-domainless protein C, a chymotrypic derivative lacking the phospholipid-binding domain, could be detected in the presence of phospholipids and Ca2+, with or without pentosan polysulphate. In a large molar excess, other phospholipid-binding entities like prothrombin fragments F1 or F1+2 could inhibit protein C activation by factor Xa, but pentosan polysulphate exerted a clear protective effect. Factor Xa irreversibly inhibited at its active centre, but not di-isopropyl phosphoro-thrombin, behaved as an inhibitor but in a more complex manner than simple Michaelis-Menten kinetics. Among several derivatives of pentosan polysulphate or of heparin which were tested, those having the higher degree of sulphation and/or molecular mass were the most efficient in enhancing the rate of activation of protein C by factor Xa in the presence of phospholipids. These results suggest that human factor Xa, at physiological concentrations, could activate human protein C in the presence of anionic phospholipids and that this activation could be potentiated by therapeutic concentrations of sulphated polysaccharides.  相似文献   

13.
R R Lobb 《Biochemistry》1988,27(7):2572-2578
Incubation of bovine brain derived acidic fibroblast growth factor (aFGF) with bovine or human thrombin, 0.5 NIH unit/mL, for 24 h at 37 degrees C results in cleavage of the mitogen, generating a 14-kilodalton fragment which has significantly reduced affinity for immobilized heparin as compared to aFGF, and is at least 50-fold less potent at stimulating mitogenesis. In addition, an 18 amino acid peptide, aFGF(123-140), is generated, identifying one of the thrombin cleavage sites as the Arg-122/Thr-123 bond. The peptide, aFGF(123-140), is neither mitogenic itself nor an inhibitor of the mitogenic activity of aFGF. The cleavage of aFGF by thrombin is inhibited by heparin (50 micrograms/mL) and is completely blocked by the irreversible thrombin inhibitors D-Phe-Pro-Arg chloromethyl ketone and hirudin. Incubation of aFGF with 50 units/mL thrombin at 37 degrees C results in rapid cleavage of the mitogen into several fragments. In contrast, incubation of bovine brain derived basic fibroblast growth factor with 1 unit/mL thrombin for 24 h, or 50 units/mL thrombin for 6 h, does not result in significant cleavage of mitogen. The results show that the C-terminal region of aFGF is of functional importance in both mitogenesis and heparin binding. Most importantly, a novel role for anionic heparin-binding growth factors and their fragments is indicated in physiologic and pathologic situations associated with thrombin generation.  相似文献   

14.
The endothelial cell surface provides a receptor for thrombin-designated thrombomodulin (TM) which regulates thrombin formation and the activity of the enzyme at the vessel wall surface by serving as a potent cofactor for the activation of protein C by thrombin. Heparin-like structures of the vessel wall have been proposed as another regulatory mechanism catalyzing the inhibition of thrombin by antithrombin III. In the present study, the interaction of antithrombin III with the thrombin-TM complex and its interference with heparin and polycations were investigated by using human components and TM isolated from the microvasculature of rabbit lung. Purified TM bound thrombin and acted as a cofactor for protein C activation. The addition of heparin (0.5 unit/mL) to the reaction mixture interfered neither with the binding of thrombin to TM nor with the activation of protein C. However, the polycations protamine (1 unit/mL) as well as polybrene (0.1 mg/mL) affected the thrombin-TM interaction. This was documented by an increase in the Michaelis constant from 8.3 microM for thrombin alone to 19.5 microM for thrombin-TM with the chromogenic substrate compound S-2238 in the presence of 1 unit/mL protamine. When the inhibition of thrombin by antithrombin III was determined, the second-order rate constant k2 = 8.4 X 10(3) M-1 s-1 increased about 8-fold in the presence of TM, implying an accelerative function of TM in this reaction. Although purified TM did not bind to antithrombin III-Sepharose, suggesting the absence of heparin-like structures within the receptor molecule, protamine reversed the accelerative effect of TM in the inhibition reaction.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
The molecular interactions between components of the heparin-catalyzed antithrombin III/thrombin reaction were investigated by light scattering. When heparin was added to antithrombin III, the molecular weight increased to a maximum and then decreased to that of a 1:1 (antithrombin III X heparin) complex. The initial molecular weights at low heparin to antithrombin III ratios were consistent with the formation of a 2:1 (antithrombin III X heparin) complex in which only one antithrombin III molecule had undergone the conformational change measured by protein fluorescence enhancement. The peak molecular weight never reached that of a complete 2:1 complex. This behavior was observed for bovine and human antithrombin III in the presence of both unfractionated heparin and high molecular weight-high affinity heparin. Pentosane polysulfate also caused some multiple associations. Bovine antithrombin III and thrombin formed a 1:1 complex that underwent further aggregation within minutes, while the human proteins did not aggregate on this time scale after forming the 1:1 complex. In the presence of stoichiometric amounts of heparin, the bovine proteins formed an initial complex of Mr = 230,000 (corresponding to a dimer of heparin-antithrombin III-thrombin) which underwent further aggregation. The human proteins, however, formed a 1:1 (antithrombin III X thrombin) initial complex in the presence of heparin, followed by aggregation. These interactions of thrombin and antithrombin with heparin suggest complex interactions that could relate to heparin function.  相似文献   

16.
17.
The binding of 125I-labeled thrombin to rat peritoneal macrophages isolated 20 h after the ip injection of thioglycollate broth or lipopolysaccharide decreased to 20% of the value found in resident macrophages due to a decrease in the number of receptors. The binding returned to normal values within a week after the injection. The decline parallelled more or less the Vmax for the 5'-nucleotidase activity. This decrease in the binding of thrombin could not be explained by an immigration of monocytes into the peritoneal cavity, since the binding of 125I-labeled alpha 2-macroglobulin-trypsin complex increased 4.5-fold in the same cell population due to an increase in the number of receptors, and blood monocytes do not bind alpha 2-macroglobulin-trypsin complex. The increase in the binding of alpha 2-macroglobulin-protease complex parallelled an increase in the incorporation of glucosamine, although the latter did not increase to the same extent. Engulfment of plasma membrane after phagocytosis did not result in a decreased binding of thrombin, but preincubation at 37 degrees C with concanavalin A caused a minor reduction in the binding. There was a positive correlation between the binding of alpha 2-macroglobulin-trypsin complex and the fraction of polymorphonuclear leukocytes in the peritoneal exudate and a negative correlation between the binding of thrombin and the fraction of polymorphonuclear leukocytes in the exudate, when the inflammation was induced by a milder stimulus, sterile NaCl, indicating a common signal for the polymorphonuclear leukocyte chemotaxis and the macrophage differentiation.  相似文献   

18.
Heparin activates the primary serpin inhibitor of blood clotting proteinases, antithrombin, both by an allosteric conformational change mechanism that specifically enhances factor Xa inactivation and by a ternary complex bridging mechanism that promotes the inactivation of thrombin and other target proteinases. To determine whether the factor Xa specificity of allosterically activated antithrombin is encoded in the reactive center loop sequence, we attempted to switch this specificity by mutating the P6-P3' proteinase binding sequence excluding P1-P1' to a more optimal thrombin recognition sequence. Evaluation of 12 such antithrombin variants showed that the thrombin specificity of the serpin allosterically activated by a heparin pentasaccharide could be enhanced as much as 55-fold by changing P3, P2, and P2' residues to a consensus thrombin recognition sequence. However, at most 9-fold of the enhanced thrombin specificity was due to allosteric activation, the remainder being realized without activation. Moreover, thrombin specificity enhancements were attenuated to at most 5-fold with a bridging heparin activator. Surprisingly, none of the reactive center loop mutations greatly affected the factor Xa specificity of the unactivated serpin or the several hundred-fold enhancement in factor Xa specificity due to activation by pentasaccharide or bridging heparins. Together, these results suggest that the specificity of both native and heparin-activated antithrombin for thrombin and factor Xa is only weakly dependent on the P6-P3' residues flanking the primary P1-P1' recognition site in the serpin-reactive center loop and that heparin enhances serpin specificity for both enzymes through secondary interaction sites outside the P6-P3' region, which involve a bridging site on heparin in the case of thrombin and a previously unrecognized exosite on antithrombin in the case of factor Xa.  相似文献   

19.

Background

The aim is to compare the structures, anticoagulant and antithrombotic activities of two fucosylated chondroitin sulfates isolated from sea cucumbers Isostichopus badionotus (fCS-Ib) and Pearsonothuria graeffei (fCS-Pg), which were known to have different sulfation patterns on the fucose branches.

Methods

The structures of fCSs were identified using 2D NMR. Anticoagulant activities were measured by activated partial thromboplastin time (APTT) and thrombin time (TT), and inhibition of factors IIa, Xa and XIIa was assessed in vitro. Antithrombotic activity was determined ex vivo by measuring the length and weight of the thrombus generated.

Results

The two fCSs had identical chondroitin sulfate E backbones and similar fucose branches, but different sulfation patterns of the fucose branches. The fucose branch in fCS-Ib was mainly 2,4-O-sulfated whereas that in fCS-Pg was mainly 3,4-O-sulfated. In vitro assay indicated that fCS-Pg possessed much lower potency than fCS-Ib in prolonging APTT/TT and in inhibition of thrombin. However, they both exhibited similar inhibitory effects on factor X activation by intrinsic tenase complex, and on thrombus generation. Furthermore, both fCSs significantly activated factor XII, which has been proved to be associated with adverse clinical events associated with heparin contaminated by oversulfated chondroitin sulfate.

Conclusion

The 2,4-O-sulfated fucose branch is the key structural factor of fCSs for prolonged APTT/TT and inhibition of thrombin, whereas the inhibitory effect of fCSs on factor X, XII activation and thrombus generation was attributed to the overall structure of fCS polysaccharide.

General importance

Both fCSs have well defined structures and can be readily quality-controlled, and therefore may be potential alternatives for heparin as anticoagulant and antithrombotic drugs.  相似文献   

20.
The serine proteinase inhibitor tissue factor pathway inhibitor-2 (TFPI-2) inhibits the tissue factor-factor VIIa complex and thereby impairs factor Xa and subsequently thrombin generation. Here we show that thrombin itself up-regulates TFPI-2 mRNA and protein expression in human liver myofibroblasts, a cell type shown to express high levels of TFPI-2 (Neaud, V., Hisaka, T., Monvoisin, A., Bedin, C., Balabaud, C., Foster, D. C., Desmoulière, A., Kisiel, W., and Rosenbaum, J. (2000) J. Biol. Chem. 275, 35565-35569). This effect required thrombin catalytic activity, as shown by its abolition with hirudin. Although the thrombin effect could be mimicked by agonists of both protease-activated receptor (PAR)-1 and PAR-4, it was largely blocked by a PAR-1 blocking antibody. Transactivation of the epidermal growth factor (EGF) receptor has been reported as a common event in thrombin signaling. However, thrombin did not detectably transactivate the EGF receptor in liver myofibroblasts, and blocking the EGF receptor did not affect TFPI-2 induction. On the other hand, thrombin increased the expression of cyclooxygenase-2 (COX-2) mRNA via a MAPK-dependent pathway, and a specific COX-2 inhibitor abolished the effect of thrombin on TFPI-2 expression. Thus, thrombin, through PAR-1 signaling, up-regulates the synthesis of TFPI-2 via a MAPK/COX-2-dependent pathway. The up-regulation of TFPI-2 expression by thrombin could in turn down-regulate thrombin generation and contribute to limit blood coagulation.  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