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 46 毫秒
1.
Y Sakata  Y Eguchi  J Mimuro  M Matsuda  Y Sumi 《Blood》1989,74(8):2692-2697
A monoclonal antibody (MoAb) to alpha 2-plasmin inhibitor designated JTPI-1 inhibited antiplasmin activity by interfering with formation of alpha 2-plasmin inhibitor (alpha 2-PI)-plasmin complex. With this MoAb, we observed plasma clot lysis in vitro and evaluated the potential of JTPI-1 to serve as a new therapeutic agent for thrombolysis. After adding 125I-labeled fibrinogen to plasma, clots were made by adding thrombin and calcium and were then resuspended in normal plasma containing various concentrations of JTPI-1. The presence of JTPI-1 enhanced release of the soluble 125I-labeled fibrin degradation fragment from the clots in a dose-dependent manner. With tissue plasminogen activator (t-PA)-depleted plasma, we showed that induction of clot lysis by JTPI-1 was dependent on fibrin-bound endogenous t-PA. Regulation of fibrinolysis initiated on the fibrin surface by fibrin-bound t-PA and plasminogen is mediated by alpha 2-PI cross-linked to fibrin by activated factor XIII. JTPI-1 bound to this cross-linked alpha 2-PI neutralized its activity and induced partial digestion of fibrin by plasmin. This resulted in additional binding of Glu-plasminogen to fibrin during the incubation. When 1.2 mumol/L JTPI-1 and 5 U/mL exogenous t-PA were present in the suspending plasma, the rate of clot lysis was essentially the same as that induced by 60 U/mL exogenous t-PA alone. These results suggest that JTPI-1 may be useful in reducing the amount of t-PA administered for thrombolytic therapy.  相似文献   

2.
Summary . The effect of fibrin crosslinking on the lysis of plasma clots was investigated with plasma from a patient congenitally deficient in plasma factor XIII (fibrin stabilizing factor). The thrombin-activated plasma factor XIII was found to render clots more resistant to fibrinolysis when urokinase (UK) was used to induce plasminogen activation. Incorporation of UK in the clot by addition to plasma immediately before clotting resulted in a log-log relationship when lysis time was plotted against UK concentration, with greater differences between normal and factor-XIII deficient clots at lower UK concentrations. Addition of UK to the clot externally, after preincubation of the clot, gave linear plots on rectangular coordinates when either plasma or euglobulin fraction was used; and the difference in lysis time between factor-XIII deficient and normal clots was nearly constant over the range of UK concentrations tested. When the fluorescent amine, dansylcadaverine, was used to measure factor-XIII activity quantitatively, plasma clot lysis times were found to be directly proportional to amine-incorporating activity over a wide range of activities at low UK concentrations. The range of proportionality was reduced when UK concentration was increased. Addition of purified factor XIII to the patient's plasma restored the resistance of these clots to within the normal range.  相似文献   

3.
Characterization of ultrasound-potentiated fibrinolysis in vitro   总被引:15,自引:0,他引:15  
Blinc  A; Francis  CW; Trudnowski  JL; Carstensen  EL 《Blood》1993,81(10):2636-2643
We have characterized the effects of ultrasound on fibrinolysis in vitro to investigate the mechanism of ultrasonic potentiation of fibrinolysis and to identify potentially useful ultrasound parameters for therapeutic application. Radiolabeled clots in thin walled tubes were exposed to ultrasound fields in a water bath at 37 degrees C, and lysis was measured by solubilization of radiolabel. Ultrasound accelerated lysis of plasma, whole blood, and purified fibrin clots mediated by recombinant tissue-type plasminogen activator (rt-PA), urokinase, or streptokinase, but ultrasound by itself caused no clot solubilization. The degree of ultrasonic potentiation was dependent on plasminogen activator concentration, increasing from 2.2-fold at a streptokinase concentration of 75 U/mL to 5.5-fold at 250 U/mL in a 1 MHz ultrasound field at 4 W/cm2. Ultrasound exposure resulted in heating due to absorption by the plastic tube, but the temperature increase was insufficient to account for the increase in clot lysis rate, indicating that the primary effect was nonthermal. Ultrasound did not accelerate hydrolysis of a peptide substrate by rt-PA and did not alter the rate of plasmic degradation of fibrinogen, indicating that the augmentation of enzymatic fibrinolysis required the presence of a fibrin gel. The acceleration of fibrinolysis by ultrasound was greater at higher intensities and duty cycles and was maximum at frequencies between 1 and 2.2 MHz, but decreased at 3.4 MHz. These findings suggest that ultrasound accelerates enzymatic fibrinolysis by increasing transport of reactants through a cavitation-related mechanism.  相似文献   

4.
Connaghan  DG; Francis  CW; Lane  DA; Marder  VJ 《Blood》1985,65(3):589-597
A new method is described for identifying low concentrations of circulating derivatives of fibrinogen and fibrin, even when present in heterogeneous mixtures. This technique is applicable to plasma and serum and uses electrophoresis in 2% agarose in the presence of sodium dodecyl sulfate (SDS) followed by immunological identification of separated derivatives, using radiolabeled antifibrinogen antiserum and autoradiography. Unique electrophoretic patterns distinguish plasmic derivatives of crosslinked fibrin from those of fibrinogen and also identify crosslinked fibrin polymers produced by the combined action of thrombin and factor XIII on fibrinogen. The assay is sensitive to a concentration of 0.1 micrograms/mL of fibrinogen in serum or plasma. Fibrin polymers, plasmic degradation products of fibrinogen, and plasmic degradation products of crosslinked fibrin were detected in the plasma or serum of a patient with disseminated intravascular coagulation. Plasmic derivatives of both fibrinogen and crosslinked fibrin appeared in serum in the course of fibrinolytic therapy for pulmonary embolism, whereas during acute myocardial infarction a marked increase in the proportion of fibrin polymers in plasma was found in comparison with normal controls. Thus, the procedure can distinguish between the simultaneous processes of fibrin polymer formation, fibrinogenolysis, and fibrinolysis, and is sufficiently sensitive to detect relevant quantities of derivatives in pathologic conditions.  相似文献   

5.
Silence  K; Collen  D; Lijnen  HR 《Blood》1993,82(4):1175-1183
The effects of alpha 2-antiplasmin and fibrin on the activation of plasminogen by recombinant staphylokinase (STAR) were studied in an effort to elucidate further the molecular basis of the fibrin- specificity of this fibrinolytic agent. In purified systems consisting of 1.5 mumol/L intact or low-M(r) plasminogen and 3 mumol/L alpha 2- antiplasmin, at 37 degrees C and in the absence of fibrin, STAR did not induce plasminogen activation and plasmin-alpha 2-antiplasmin complex (PAP) formation. Addition of a purified fibrin clot (30% vol at a concentration of 3 mg/mL) to mixtures containing intact plasminogen caused approximately 40% plasminogen activation within 2 hours, whereas in mixtures containing low-M(r) plasminogen, no activation was observed. In contrast, 10 nmol/L streptokinase (SK) induced 74% to 100% plasminogen activation within 2 hours in mixtures containing either intact or low-M(r) plasminogen, in both the absence and the presence of fibrin. In citrated human plasma in the absence of fibrin, 30 nmol/L STAR did not induce measurable plasminogen activation and PAP formation (< 1.5% within 2 hours), whereas addition of a plasma clot (12% vol) resulted in complete clot lysis and conversion of 19% +/- 8% of the plasminogen to PAP within 2 hours. Addition of a second plasma clot produced 23% +/- 2% additional plasminogen activation. Equipotent concentrations for plasma clot lysis of SK (100 nmol/L) induced 54% +/- 11% plasminogen activation in the absence and 49% +/- 16% in the presence of fibrin. Addition of 50 mmol/L 6-aminohexanoic acid (6-AHA) abolished the effect of fibrin on plasminogen activation with STAR, but not on activation with SK. In alpha 2-antiplasmin-depleted human plasma in the absence of fibrin, 30 nmol/L STAR did not induce fibrinogen breakdown (> 90% residual fibrinogen after 6 hours), whereas 30 nmol/L preformed plasmin-STAR complex induced extensive fibrinogen degradation (70% within 20 minutes). Thus, in the absence of fibrin, alpha 2- antiplasmin inhibits the activation of plasminogen by STAR, by preventing generation of active plasmin-STAR complex. Fibrin stimulates plasminogen activation by STAR via mechanisms involving the lysine- binding sites of plasminogen, probably by facilitating the generation of plasmin-STAR complex and by delaying its inhibition at the clot surface.  相似文献   

6.
A murine monoclonal antibody (MA-12E6A8) was raised against human urokinase-type plasminogen activator (u-PA), which, in an enzyme-linked immunosorbent assay (ELISA), reacted 15,000-fold better with recombinant two-chain u-PA (rtcu-PA) than with recombinant single-chain u-PA (rscu-PA). The antibody had no effect on the activity of rtcu-PA or on its inhibition by a chloromethylketone, but reduced the inhibition of rtcu-PA by recombinant plasminogen activator inhibitor-1 (rPAI-1) at least 10-fold. The dissociation constant of the rtcu-PA/MA-12E6A8 complex was 7 nmol/L. An ELISA was developed using MA-12E6A8 as capture antibody and a horseradish peroxidase conjugated u-PA specific antibody for tagging. It recognized free and active site blocked rtcu-PA but not rtcu-PA in complex with rPAI-1 or with alpha 2-antiplasmin. This ELISA was used to monitor the generation of rtcu-PA during fibrin clot lysis with rscu-PA in human plasma. Addition of 5 micrograms/mL rscu-PA to 3 mL plasma containing a 0.2 mL 125I-fibrin labeled plasma clot caused 50% clot lysis in 62 +/- 13 minutes (mean +/- SD, n = 6), at which time 99 +/- 28 ng/mL rtcu-PA was detected but no fibrinogen breakdown had occurred. Fifty percent fibrinogen breakdown did occur only when rtcu-PA had reached a level of 1,000 +/- 270 ng/mL (at 150 +/- 21 minutes). rscu-PA, 2 micrograms/mL, induced 50% clot lysis in 160 +/- 41 minutes (n = 6); no fibrinogen degradation occurred within 4 hours and rtcu-PA levels did not exceed 80 ng/mL. In the absence of a fibrin clot, 5 micrograms/mL rscu-PA added to human plasma did not result in significant generation of rtcu-PA (less than 50 ng/mL after 4 hours) and no fibrinogen degradation was observed. These results indicate that clot lysis with rscu-PA in a plasma milieu does not require extensive systemic conversion of rscu-PA to rtcu-PA, and that fibrinogen degradation occurs secondarily to systemic conversion of rscu-PA to rtcu-PA.  相似文献   

7.
Sakata  Y; Mimuro  J; Aoki  N 《Blood》1984,63(6):1393-1401
In spontaneous fibrinolysis of an alpha 2-plasmin inhibitor-deficient plasma clot or tissue-type plasminogen activator-induced fibrinolysis in a purified system without alpha 2-plasmin inhibitor, the lysis was faster when factor XIII-mediated crosslinking of fibrin to fibrin did not occur. During the initial period, the binding of plasminogen to fibrin steadily increased with incubation time. The initial level and subsequent increase of the binding, which may be critical for the subsequent development of fibrinolysis, were more remarkable when fibrin was not crosslinked. The amount of glu- or lys-plasminogen bound to noncrosslinked fibrin was around 4 or 1.5 times larger than the amount of the respective plasminogen bound to crosslinked fibrin. Plasmin was also found to be bound to noncrosslinked fibrin twice as much as the amount bound to crosslinked fibrin. Structural changes induced by crosslinking of fibrin alpha-chain may reduce either the affinity or the number of available complementary sites to lysine binding sites of plasmin(ogen), thereby decreasing the binding of plasmin(ogen) to fibrin. These results suggest that an increased affinity of noncrosslinked fibrin for plasmin(ogen) is contributory to the accelerated fibrinolysis observed in factor XIII deficiency, in addition to an absence of crosslinking of alpha 2-plasmin inhibitor to fibrin.  相似文献   

8.
A fibrin glue preparation has been obtained from pooled human plasma using a procedure which includes a solvent-detergent (SD) treatment to inactivate lipid-enveloped viruses. The SD treatment inactivated greater than or equal to 5.5 log10 of HIV in less than 45 min, and greater than or equal to 5 log10 and greater than or equal to 6.5 log10 of VSV and Sindbis virus, respectively, in less than 2 h. The product was found to contain high quantities of fibrinogen (116 +/- 2.49 g/l; n = 12), factor XIII (35 +/- 2.88 U/ml) and von Willebrand factor (23 +/- 1.9 U/ml ristocetin cofactor activity), and relatively low levels of fibronectin (5.9 +/- 0.51 g/l). Plasminogen, the precursor of plasmin, which may play a negative role by decreasing the resistance of the fibrin clot, was at only 0.03 g/l. Cellulose acetate electrophoresis showed 95% gamma-proteins and 5% alpha-2-beta proteins. Sodium dodecyl sulfate polyacrylamide gel electrophoresis under reducing conditions detected three main protein bands with apparent molecular weights of 65, 56 and 47 kilodaltons, probably corresponding to the alpha, beta, and gamma fibrinogen subunits. Other characteristics of the product included (1) high clottability of fibrinogen (over 85%); (2) absence of low molecular weight fibrin degradation products; (3) rapid solubilization at room temperature (less than 10 min); (4) high tensile strength (202 +/- 27 g/cm2 after 2 h of application), and (5) high elasticity of the fibrin clot. In addition, scanning electron microscopy revealed a highly organized structure showing tridimensional arrangement of the fibrin fibers. SD treated fibrin glue should efficiently replace autologous fibrinogen or cryoprecipitate preparations for surgical application.  相似文献   

9.
Kimura  S; Tamaki  T; Aoki  N 《Blood》1985,66(1):157-160
When blood plasma containing the NH2-terminal 12-residue peptide (N- peptide) of alpha 2-plasmin inhibitor (alpha 2PI; alpha 2-antiplasmin) was clotted in the presence of calcium ions, the N-peptide and alpha 2PI were cross-linked to fibrin by activated coagulation factor XIII. The amount of N-peptide cross-linked to fibrin was proportional to the concentration of N-peptide present in plasma. On the other hand, the amount of alpha 2PI cross-linked to fibrin was decreased by the presence of N-peptide, and the decrease was in reverse relationship to the increase of cross-linking of N-peptide. Spontaneous fibrinolysis or fibrinolysis induced by tissue plasminogen activator was accelerated by the presence of N-peptide, and the acceleration was dependent on the concentrations of N-peptide and directly proportional to inhibition of alpha 2PI cross-linking exerted by N-peptide. The acceleration was more pronounced when the clot was compacted by platelet-mediated clot retraction or by a squeeze. Fibrinolysis of an alpha 2PI-deficient or a factor XIII-deficient plasma clot was not accelerated by N-peptide. These findings were substantiated in a purified system and support the previous proposal that alpha 2PI is cross-linked to fibrin at the glutamine residue that is next to the NH2-terminus of alpha 2PI, and this factor XIII-mediated cross-linking of alpha 2PI is significant in inhibition of physiologically occurring endogenous fibrinolysis.  相似文献   

10.
Although heparin is currently used in concomitance with thrombolytic agents to improve their efficacy, its effect on fibrinolysis is controversial. We have evaluated the sensitivity to t-PA-induced lysis of clots prepared from plasma preincubated in vitro with therapeutic concentrations of heparin. The extent of t-PA-induced lysis was significantly increased by preincubation of plasma with 0.5 and 1.0 U/ml heparin. The concentration of t-PA required to give similar lysis rates were reduced by up to five times after adding 1.0 U/ml heparin to plasma prior to clot formation. Heparin added to the t-PA-containing medium after clot formation did not exert any significant effect. The effect of heparin was not mediated by the inhibition of thrombin as preincubation of plasma with hirudin did not modify clot sensitivity to t-PA. We also found that heparin significantly modified fibrin assembly and clot structure as assessed by a turbidimetric assay. Pre-incubation of fibrinogen with heparin caused an increase in the speed of fibrin fibre polymerization and in the turbidity of the final fibrin gel; changes known to be associated with the formation of thicker fibrin fibres. Thus the effect of heparin on clot sensitivity to lysis appears to be due to an increased permeability of these clots to fibrinolytic components. This may contribute to the antithrombotic activity and to the haemorrhagic risk of heparin. These findings could be particularly important for clinical thrombolysis.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
Since intravascular and endoparietal fibrin deposition is thought to be involved in the development of atherosclerosis, we measured factor XIII activity and its subunit 'a' and 'b' concentrations against a background of other haemostasis parameters in diabetics with angiopathy and in 2 control groups (healthy subjects and diabetics without vascular complications). Diabetics with angiopathy revealed a significant increase of factor XIII activity as well as its subunit concentrations. They also had significantly elevated anti-thrombin III, alpha 2 macroglobulin, alpha 1 antitrypsin, C1 inhibitor, fibrinogen, FDP concentrations and prolongation of euglobulin lysis time. The highest factor XIII levels were found in diabetics with renal failure. We suppose that increased factor XIII level and other observed changes of haemostasis in patients with diabetic angiopathy might promote intravascular and endoparietal fibrin deposition and contribute to the development of atherosclerotic complications of diabetes.  相似文献   

12.
Greenberg  CS; Achyuthan  KE; Fenton  JW d 《Blood》1987,69(3):867-871
Fibrin polymers (des A,B fibrinogen) reduced the concentration of alpha- thrombin required for 50% activation of plasma factor XIII (a2b2 tetramer) by approximately 100-fold. In the presence of fibrin, the amount of gamma-thrombin required for activation was not affected. Catalytically inactive i-Pr2P- and D-Phe-Pro-Arg-CH2-alpha-thrombin were found to inhibit over 95% of the activation by alpha-thrombin in the presence of fibrin. Unlike plasma factor XIII, the concentration of alpha-thrombin required for 50% activation of platelet factor XIII (a2 dimer) was lower, and the activation was not enhanced by fibrin. However, when the a2 platelet factor XIII was incubated with purified b- chains, the alpha- and gamma-thrombin concentrations required for activation increased tenfold and reached levels similar to those required for activation of the plasma factor XIII. When fibrin was present, the alpha-thrombin concentrations needed for activation of the a2b2 complexes were reduced, and the presence of fibrin had no effect on gamma-thrombin cleavage of the a2b2 complexes. Therefore, the b- chains must inhibit a-chain cleavage by alpha-thrombin in the absence of fibrin. These results imply that the formation of a cocomplex involving alpha-thrombin, fibrin, and plasma factor XIII causes some conformational change in plasma factor XIII such that the b-chains no longer inhibit cleavage of the a-chains.  相似文献   

13.
S ummary . An acquired haemorrhagic syndrome is described in which the failure of haemostasis is due to impaired crosslinking of fibrin. This defect, demonstrated by both monochloroacetic acid clot-solubility test and disc-gel electrophoretic analysis of individual fibrin chains, existed despite normal fibrin stabilizing factor (factor XIII) levels in the patient's plasma. Activation of the factor by thrombin, measured by amine incorporation into casein, was normal.
The unique molecular defect in this patient could be fully accounted for by the presence of an inhibitory IgG antibody against the crosslinking sites in fibrinogen-fibrin. No alteration could be seen in either the rate or in the extent of the reversible aggregation of fibrin. However, large amounts of cryoprecipitate were found in the patient's plasma. This phenomenon could be reproduced in vitro by incubating the chromatographically purified IgG of the patient with normal fibrinogen at 4°C.
It is conceivable that the antibody in the patient could have arisen in response to the modification of normal fibrinogen by isonicotinic acid hydrazide which the patient had ingested for 8 yr.  相似文献   

14.
Regulation of plasma factor XIII binding to fibrin in vitro   总被引:2,自引:2,他引:2  
C S Greenberg  J V Dobson  C C Miraglia 《Blood》1985,66(5):1028-1034
The binding of plasma factor XIII to fibrinogen or fibrin that has been chemically or enzymatically induced to polymerize was studied. Factor XIII binding was assayed using a 3H-putrescine incorporation assay and an 125I-plasma factor XIII binding assay. More than 80% of the native and radiolabeled plasma factor XIII was bound to fibrin I formed by reptilase in EDTA, citrate, or heparin anticoagulated plasma. Plasma factor XIII and 125I-factor XIII was bound (89.6% to 92.5%) to fibrin II formed by thrombin in either citrate or EDTA anticoagulated plasma. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) of 125I-plasma factor XIII bound to fibrin I or fibrin II formed by reptilase or thrombin in the presence of EDTA demonstrated the b2-subunit remained bound to the a-chains or thrombin-cleaved a-chains. In the presence of calcium chloride and thrombin, the b2-subunit dissociated and factor XIIIa was bound. Protamine sulfate caused fibrinogen polymerization in the absence of divalent cations and reduced both plasma factor XIII and immunologic fibrinogen levels. Fibrinogen polymerized by protamine sulfate bound plasma factor XIII and the a2-subunit of 125I-platelet factor XIII. Plasma factor XIII was also bound to sonicated non-cross-linked fibrin II in either normal plasma or afibrinogenemic plasma. Plasma levels of several coagulation proteins were unchanged after the addition of reptilase, protamine sulfate, or sonicated fibrin to plasma. These results demonstrate that a specific binding site for the a2-subunit of plasma factor XIII is present on polymerized fibrinogen, fibrin I, and fibrin II. Furthermore, the presence of divalent cations, thrombin-cleavage of plasma factor XIII, and release of fibrinopeptides A or B are not required for plasma factor XIII binding to polymerized fibrinogen and fibrin.  相似文献   

15.
A mutant of recombinant tissue-type plasminogen activator (rt-PA), obtained by deletion of residues Lys296 to Gly302 [rt-PA del(K296-G302)], was previously shown to be resistant to inhibition by plasminogen activator inhibitor-1 (PAI-1) (Madison et al, Nature 339:721, 1989). This mutant was obtained by expression of its cDNA in Chinese hamster ovary cells and purification to homogeneity from conditioned cell culture medium. It was obtained as a single chain molecule with amidolytic activity, specific fibrinolytic activity, and binding to fibrin and lysine, which were comparable or somewhat lower than those of wild-type rt-PA obtained in the same expression system. The plasminogen-activating potential of rt-PA del(K296-G302) in the presence of CNBr-digested fibrinogen was about twofold lower than that of wild-type rt-PA. The inhibition rate of rt-PA del(K296-G302) by recombinant PAI-1 (rPAI-1) was more than 500-fold lower than that of wild-type rt-PA. In a human plasma milieu in vitro, rt-PA del(K296-G302) induced dose-dependent lysis of a 125I-fibrin-labeled plasma clot; equi-effective concentrations (causing 50% clot lysis in 2 hours) were 0.28 micrograms/mL and 0.36 micrograms/mL for mutant and wild-type rt-PA, respectively. In this system, addition of rPAI-1 to the plasma resulted in a concentration-dependent reduction of the fibrinolytic potency of rt-PA del(K296-G302) and of rt-PA; a 50% reduction required 2.4 micrograms/mL and 0.15 micrograms/mL rPAI-1, respectively. Continuous infusion of mutant or wild-type rt-PA over 60 minutes in hamsters with a 125I-labeled plasma clot in the pulmonary artery resulted in dose-dependent clot lysis, with a thrombolytic potency (percent clot lysis per milligram of compound administered per kilogram of body weight) and a specific thrombolytic activity (percent clot lysis per microgram per milliliter steady state rt-PA-related antigen level in plasma) that were not significantly different. Bolus injection in hamsters of 1 mg/kg rPAI-1 followed by bolus injection of 1 mg/kg rt-PA del(K296-G302) or wild-type rt-PA resulted in neutralization of the thrombolytic potency of wild-type rt-PA, while the mutant retained approximately half of its thrombolytic potency. These results indicate that rt-PA del(K296-G302), with a known resistance to inhibition by rPAI-1 in purified systems, maintains this property both in a plasma milieu in vitro and in an experimental animal model of thrombolysis in vivo.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

16.
The S-carboxymethyl derivative chains of crosslinked and noncrosslinked fibrins were prepared from purified human fibrinogen. For crosslinked fibrin, fibrinogen was clotted with thrombin in the presence of calcium and purified human factor XIII. For noncrosslinked fibrin, ethylenediaminetetraacetate was substituted for factor XIII and calcium. After reduction with dithiothrcitol and alkylation with tritiated iodoacetic acid, the derivative chains were separated on carboxymethyl cellulose in a sodium acetate-pH gradient that contained 8 M urea. Purity of the separated chains was determined by polyacrylamide gel electrophoresis at acid and at neutral pH. The derivative chains of noncrosslinked fibrin were eluted from carboxymethyl cellulose in the order: gamma-chain, beta-chain, and alpha-chain. Each of the purified derivative chains was characterized and identified by amino-terminal aminoacid analysis, aminoacid composition, tryptic peptide mapping, and molecular weight estimation by polyacrylamide gel electrophoresis in sodium dodecyl sulfate. In like manner, the derivative components of crosslinked fibrin were eluted from carboxymethyl cellulose in the order: gamma-gamma-dimer, beta-chain, and alpha-polymer. Application of the same analytical criteria and comparision with the derivatives of noncrosslinked fibrin confirmed the identity of these components. These data provide conclusive evidence that crosslinking of human fibrin involves formation of peptide bonds between two gamma-chains to form gamma-gamma-dimer and between multiple alpha-chains to form high molecular weight polymers of alpha-chains.  相似文献   

17.
Holvoet  P; Lijnen  HR; Collen  D 《Blood》1986,67(5):1482-1487
One (MA-1C8) of 36 monoclonal antibodies obtained by fusion of P3X63- Ag8-6.5.3 myeloma cells with spleen cells of mice immunized with purified human tissue-type plasminogen activator (t-PA) blocked the activity of t-PA on fibrin plates but not on chromogenic substrates. MA- 1C8 at a concentration of 200 micrograms/mL inhibited plasma clot lysis and binding of t-PA to the clot. MA-1C8 had no influence on the activation of plasminogen by t-PA, which obeys Michaelis-Menten kinetics with Km = 105 mumol/L and kcat = 0.05 s-1; however, it abolished the influence of CNBr-digested fibrinogen on Km. These findings confirm that the stimulatory effect of fibrin on the activation of plasminogen by t-PA is mediated by binding of t-PA to fibrin and provide additional support for the kinetic model. Addition of t-PA to pooled fresh human plasma to a concentration of 5 micrograms/mL resulted in extensive fibrinogen breakdown after incubation for one hour at 37 degrees C or during storage at -20 degrees C for one day. In both instances, fibrinogen degradation was completely prevented by addition of MA-1C8 to a concentration of 200 micrograms/mL of plasma. MA-1C8 also effectively prevented in vitro fibrinogen degradation and in vitro plasminogen activation in plasma samples obtained during infusion of recombinant t-PA in patients with thromboembolic disease. Thus, MA-1C8 is a useful tool for discriminating between in vivo and in vitro fibrinolysis during thrombolytic therapy with t-PA.  相似文献   

18.
Thrombin cleaves fibrinopeptides from fibrinogen, converting it to fibrin monomer, and activates factor XIII, which catalyzes the formation of intermolecular epsilon-(gamma-glutamyl)-lysine bonds to stabilize the fibrin polymer. The formation of factor XIIIa-catalyzed fibrin polymers during clotting of plasma and purified fibrinogen in vivo was followed by a sodium dodecyl sulfate agarose gel technique, and an increase in both amount and size of gamma-chain cross-linked polymers was demonstrated before visible clot formation. Plasma from patients presenting with acute myocardial infarction showed increases in the plasma concentration of fibrin polymer and in the proportion of total fibrinogen present as polymer, as determined by a quantitative adaptation of the electrophoretic technique. The plasma concentration in patients with subendocardial or transmural myocardial infarction showed significant (p less than .005) increases to 4.0 +/- 1.0% and 3.6 +/- .8%, respectively, as compared with the concentration in normal plasma (0.8 +/- 0.1%). There was no difference in plasma concentration in samples from patients with transmural compared with those with subendocardial myocardial infarction. This study provides the first demonstration of factor XIIIa cross-linked fibrin polymers in thrombotic disease and indicates the presence of increased activity of both thrombin and factor XIIIa in patients with acute myocardial infarction.  相似文献   

19.
Lee KN  Jackson KW  Christiansen VJ  Chung KH  McKee PA 《Blood》2004,103(10):3783-3788
Human alpha2-antiplasmin (alpha2AP), also known as alpha2-plasmin inhibitor, is the major inhibitor of the proteolytic enzyme plasmin that digests fibrin. There are 2 N-terminal forms of alpha2AP that circulate in human plasma: a 464-residue protein with Met as the N-terminus, Met-alpha2AP, and a 452-residue version with Asn as the N-terminus, Asn-alpha2AP. We have discovered and purified a proteinase from human plasma that cleaves the Pro12-Asn13 bond of Met-alpha2AP to yield Asn-alpha2AP and have named it antiplasmin-cleaving enzyme (APCE). APCE is similar in primary structure and catalytic properties to membrane-bound fibroblast activation protein/seprase for which a physiologic substrate has not been clearly defined. We found that Asn-alpha2AP becomes cross-linked to fibrin by activated factor XIII approximately 13 times faster than native Met-alpha2AP during clot formation and that clot lysis rates are slowed in direct proportion to the ratio of Asn-alpha2AP to Met-alpha2AP in human plasma. We conclude that APCE cleaves Met-alpha2AP to the derivative Asn-alpha2AP, which is more efficiently incorporated into fibrin and consequently makes it strikingly resistant to plasmin digestion. APCE may represent a new target for pharmacologic inhibition, since less generation and incorporation of Asn-alpha2AP could result in a more rapid removal of fibrin by plasmin during atherogenesis, thrombosis, and inflammatory states.  相似文献   

20.
Because there was a possibility that activated factor XIII (factor XIIIa) might stabilize a platelet-fibrinogen aggregate through its crosslinking action, we have isolated plasma factor XIII, activated it, and studied the effect of factor XIIIa at a concentration of 3.3 micrograms/ml on aggregation and 125I-fibrinogen binding of rabbit platelets stimulated with 9 microM ADP. Factor XIIIa did not cause aggregation in the absence of ADP, nor did it enhance ADP-induced aggregation or substantially stabilize the platelet aggregate. The presence of factor XIIIa did not affect the amount of fibrinogen bound to platelets immediately after stimulation with ADP, but it appeared to cause a slow specific binding of 125I-fibrinogen to platelets whether or not they were stimulated with ADP. This binding, which was not inhibited by prostaglandin E1, did not lead to aggregation and was accompanied by crosslinking of fibrinogen through its A alpha and gamma chains, either to other fibrinogen molecules or to a platelet protein or proteins.  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