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.
C S Chen  S H Chou  P Thiagarajan 《Biochemistry》1988,27(16):6121-6126
The binding of fibrinogen to activated platelets leads to platelet aggregation. Fibrinogen has multiple binding sites to platelet membrane glycoprotein IIb-IIIa complex. At least two well-defined sequences in fibrinogen, Arg-Gly-Asp sequence of A alpha 95-97 and A alpha 572-574 and gamma 400-411, have been shown to interact with glycoprotein IIb-IIIa. A possible binding site on the amino-terminal end of fibrinogen to platelet glycoprotein IIb-IIIa has also been reported. In this paper the effect of synthetic peptides derived from the amino-terminal end of the B beta chain on platelet aggregation and fibrinogen binding has been examined. B beta 15-42 peptide inhibits platelet aggregation and 125I-fibrinogen binding to activated platelets in a dose-dependent manner. Since B beta 15-42 contains a previously identified fibrinogen binding site, B beta 15-18, exposed by thrombin cleavage of native fibrinogen, we also examined the effect of B beta 15-18, B beta 19-42, and B beta 1-14 (fibrinopeptide B) on platelet aggregation and fibrinogen binding. Synthetic fibrinopeptide B and B beta 15-18 had no effect on platelet aggregation and fibrinogen binding while B beta 19-42 retained the inhibitory effect. When fibrinogen is chromatographed on a column of agarose-bound B beta 15-42, a cation-dependent retention of fibrinogen on the peptide column was observed, and fibrinogen was eluted from the column by B beta 15-42 but not by B beta 1-14. Under the same conditions, platelet glycoprotein IIb-IIIa was not retained in the column. Thus, the observed inhibitory effect is due to its interaction with fibrinogen rather than to platelet glycoprotein IIb-IIIa.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
Factors which stimulate the release of von Willebrand factor (vWf) from endothelial cell Weibel-Palade bodies and which induce the expression of the leukocyte-binding adhesion molecule P-selectin (PADGEM, GMP-140, CD62) on the endothelial cell surface remain incompletely characterized. Fibrin but not fibrinogen is a potent stimulus for the release of stored von Willebrand factor from endothelial cells. Removal of fibrinopeptides A and B from fibrinogen occurs during the formation of fibrin, and the removal of fibrinopeptide B is a requirement for fibrin to induce vWf secretion. The cleavage of fibrinopeptide A by reptilase enzyme forms a fibrin gel yet it is incapable of stimulating Weibel-Palade body degranulation. As a consequence of removing fibrinopeptide B, B beta 15-42 becomes the new NH2 terminus of the beta chain of fibrin. We have shown that the peptide B beta 15-42 in solution inhibits the release of vWf stimulated by fibrin. In addition, B beta 15-42 coupled to ovalbumin supports the binding and spreading of endothelial cells, while a scrambled form of this peptide coupled to the same carrier does not. We investigated whether these determinants near the amino terminus of the beta chain of fibrin bind to a specific protein on the surface of endothelial cells. A 130-kDa protein was isolated from surface-labeled human umbilical vein endothelial cells by specific binding to B beta 15-42 immobilized on Sepharose. This glycoprotein was eluted with the B beta 15-42 peptide in solution but not with the scrambled form of this peptide. The fibrin-derived peptides B beta 19-26 and B beta 37-56-cysteine were also incapable of eluting the 130-kDa protein bound to immobilized B beta 15-42 as were the arginine-glycine-aspartic acid-serine RGDS tetrapeptide and EDTA. The 130-kDa protein is recognized neither by antibodies to the known integrins found on endothelial cells nor by antibodies to CD31 (endoCAM, PECAM-1), a member of the immunoglobulin family of receptors found on endothelial cells. The beta chain of fibrin thus contains a sequence near its amino terminus which specifically binds to what is likely a novel endothelial cell surface protein. This glycoprotein may promote endothelial cell adhesion to fibrin during the wound healing process and is a candidate for a receptor involved in fibrin-mediated release of Weibel-Palade bodies from endothelial cells.  相似文献   

3.
Employing high-performance liquid chromatography (HPLC), we have isolated and quantified the peptides that are released from the NH2-terminus of human fibrinogen B beta-chains by plasmin proteolysis. The peptides were identified by amino acid composition and by a radioimmunoassay developed for fibrinopeptide B detection. B beta 1-42 was the earliest fragment released during limited plasmin proteolysis. The level of this peptide reached a maximum and then began to decline during the course of the digestion. In addition, increasing levels of B beta 1-21 and of FPB followed the production of B beta 1-42. Using purified B beta 1-42 as a substrate, preferential cleavage was shown to occur at the 21-22 bond, with a minor cleavage at the 14-15 bond. Exhaustive digestion yielded two major components which were separated by HPLC: B beta 1-14 (FPB) and beta 22-42. The rate of cleavage at the 14-15 bond, which is the customary site of thrombin proteolysis, was not affected by the addition of hirudin indicating that this was not the result of trace contamination with thrombin. We have also examined plasmin proteolysis at the NH2-terminal region of the B beta-chains of a variety of fibrinogen derivatives and have found similar patterns of B beta 1-42 release. Using HPLC data, we have estimated the Km for plasmic cleavage of the beta 21-22 bond to be 1.8 X 10(-5) M and of the beta 14-15 bond to be 2.8 X 10(-5) M.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
Out of 29 disulfide bonds in human fibrinogen, 7 were cleaved during limited reduction under nondenaturing conditions in calcium-free buffer: 2 A alpha 442Cys-A alpha 472Cys and 2 gamma 326Cys-gamma 339Cys intrachain disulfide bonds in the carboxy-terminal ends of the A alpha- and gamma-chains and the symmetrical disulfide bonds at gamma 8Cys, gamma 9Cys, and A alpha 28Cys. We studied the loss of thrombin clottability that followed limited reduction and the increase in the susceptibility of the fibrinogen A alpha 19-A alpha 20 bond to hydrolysis by thrombin. Using differential scanning calorimetry, we show that the extent of unfolding and denaturation of specific domains following limited reduction is small. Heat absorption peaks corresponding to the melting of the major regions of compact structure give high calorimetric enthalpies, as in untreated nonreduced fibrinogen, indicating that substantial regions of native structure are still present in partially reduced fibrinogen. Thrombin releases fibrinopeptide A at an identical rate as in nonreduced fibrinogen while fibrinopeptide B release is slower. Sedimentation velocity studies show that thrombin treatment leads to complex formation; however, gelation does not occur. Amino-terminal analysis indicates that the second thrombin cleavage in the A alpha-chain at A alpha 19-A alpha 20 takes place only after fibrinopeptide A release. Thus, the loss of clottability appears to result from perturbation of carboxy-terminal polymerization sites, probably a consequence of gamma 326Cys-gamma 339Cys intrachain disulfide bond cleavage. The thrombin-treated partially reduced fibrinogen remains soluble in buffered saline and fully expresses at least one epitope, B beta 15-21, unique to fibrin. Furthermore, this nonclottable form accelerates the tissue plasminogen activator dependent conversion of plasminogen to plasmin.  相似文献   

5.
The carboxyl-terminal regions of the fibrinogen Aalpha chains (alphaC regions) form compact alphaC-domains tethered to the bulk of the molecule with flexible alphaC-connectors. It was hypothesized that in fibrinogen two alphaC-domains interact intramolecularly with each other and with the central E region preferentially through its N-termini of Bbeta chains and that removal of fibrinopeptides A and B upon fibrin assembly results in dissociation of the alphaC regions and their switch to intermolecular interactions. To test this hypothesis, we studied the interactions of the recombinant alphaC region (Aalpha221-610 fragment) and its subfragments, alphaC-connector (Aalpha221-391) and alphaC-domain (Aalpha392-610), between each other and with the recombinant (Bbeta1-66)2 and (beta15-66)2 fragments and NDSK corresponding to the fibrin(ogen) central E region, using laser tweezers-based force spectroscopy. The alphaC-domain, but not the alphaC-connector, bound to NDSK, which contains fibrinopeptides A and B, and less frequently to desA-NDSK and (Bbeta1-66)2 containing only fibrinopeptides B; it was poorly reactive with desAB-NDSK and (beta15-66)2 both lacking fibrinopeptide B. The interactions of the alphaC-domains with each other and with the alphaC-connector were also observed, although they were weaker and heterogeneous in strength. These results provide the first direct evidence for the interaction between the alphaC-domains and the central E region through fibrinopeptide B, in agreement with the hypothesis given above, and indicate that fibrinopeptide A is also involved. They also confirm the hypothesized homomeric interactions between the alphaC-domains and display their interaction with the alphaC-connectors, which may contribute to covalent cross-linking of alpha polymers in fibrin.  相似文献   

6.
To examine whether intravascular coagulation and/or decreased fibrinolysis precedes high-altitude pulmonary edema (HAPE) we examined 25 male mountaineers (median age 40 yr) at low altitude (550 m) and after 6, 18, and 42 h at an altitude of 4,559 m, which was climbed in 24 h. In 14 subjects, 2 of whom showed radiological evidence of HAPE after 42 h, symptoms of acute mountain sickness (AMS) were mild or absent. Eleven subjects suffered from AMS, six of whom developed radiologically documented HAPE after 18 or 42 h. In the absence of AMS there were no significant changes at high altitude, with the exception of a decrease in bleeding time from 246 +/- 18 to 212 +/- 13 (SE) (P less than 0.05). In AMS, partial thromboplastine time decreased from 34.2 +/- 0.8 to 31.1 +/- 0.5 s (P less than 0.001) and factor VIII procoagulant activity and von Willebrand factor antigen were increased by 57 +/- 12 and 70 +/- 13%, respectively (P less than 0.001), whereas there were no significant changes in beta-thromboglobulin (BTG), fibrinopeptide A (FPA), and fibrin fragment B beta 15-42. In subjects with HAPE, BTG, FPA, and B beta 15-42 were normal before and in beginning HAPE. Preceding HAPE, euglobulin clot lysis time declined at high compared with low altitude from 289 +/- 48 to 201 +/- 42 min without venous occlusion (VO) and from 107 +/- 36 to 86 +/- 31 min after VO (P less than 0.05).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
Multiple factors affect the thrombin-catalyzed conversion of fibrinogen to fibrin, including: fibrinopeptide (FPA and FPB) release leading to exposure of two types of polymerization domains ("A" and "B," respectively) in the central portion of the molecule, and exposure of a noncatalytic "secondary" thrombin-binding site in fibrin. Fibrinogen containing the FPA sequence but lacking the B beta 1-42 sequence ("des-(B beta 1-42)-fibrinogen"), was compared to native fibrinogen (containing both FPA and FPB) to investigate the role played by B beta 1-42 in the polymerization of alpha-fibrin (i.e. fibrin lacking FPA), to compare reptilase and thrombin cleavage of FPA from fibrinogen, and to explore the location and function of the secondary thrombin-binding site. Electron microscopy of evolving polymer structures (mu, 0.14; pH 7.4) plus turbidity measurements, showed that early thin fibril formation as well as subsequent lateral fibril associations were impaired in des-(B beta 1-42)-alpha-fibrin, thus indicating that the B beta 1-42 sequence contributes to the A polymerization site. Reptilase-activated des-(B beta 1-42)-alpha-fibrin polymerized even more slowly than thrombin-activated des-(B beta 1-42)-alpha-fibrin, differences that disappeared when repolymerization of preformed fibrin monomers was carried out. Since existing data indicate that thrombin releases FPA in a concerted manner, resulting in relatively rapid evolution of fully functional divalent alpha-fibrin monomers, it can be inferred that delayed fibrin assembly of reptilase fibrin is due to slower formation of divalent alpha-fibrin monomers. Thrombin-activated des-(B beta 1-42)-alpha-fibrin polymerized more rapidly at low ionic strength (mu, 0.04) than did native alpha,beta-fibrin, a reversal of their behavior at physiological ionic strength (mu, 0.14). Concomitant measurement of FPA release revealed modest slowing of release at low ionic strength from des-(B beta 1-42)-fibrinogen (t1/2, 36.5 versus 21.5 min) and marked slowing from native fibrinogen (t1/2, 138 versus 22.2 min). This behavior correlated with increased thrombin binding to native alpha,beta-fibrin at low ionic strength, coupled with weak thrombin binding to des-(B beta 1-42)-alpha-fibrin, and indicates that secondary thrombin binding plays an important role in regulating thrombin diffusion and catalytic activity. Des-(B beta 1-42)-fibrinogen lacks or has a markedly defective secondary thrombin-binding site, from which we conclude that the B beta 15-42 sequence in fibrin plays a major role in forming or providing this site.  相似文献   

8.
During cleavage of fibrinogen by thrombin, fibrinopeptide A (FpA) release precedes fibrinopeptide B (FpB) release. To examine the basis for this ordered release, we synthesized A'beta fibrinogen, replacing FpB with a fibrinopeptide A-like peptide, FpA' (G14V). Analyses of fibrinopeptide release from A'beta fibrinogen showed that FpA release and FpA' release were similar; the release of either peptide followed simple first-order kinetics. Specificity constants for FpA and FpA' were similar, demonstrating that these peptides are equally competitive substrates for thrombin. In the presence of Gly-Pro-Arg-Pro, an inhibitor of fibrin polymerization, the rate of FpB release from normal fibrinogen was reduced 3-fold, consistent with previous data; in contrast, the rate of FpA' release from A'beta fibrinogen was unaffected. Thus, with A'beta fibrinogen, fibrinopeptide release from the beta chain is similar to fibrinopeptide release from the alpha chain. We conclude that the ordered release of fibrinopeptides is dictated by the specificity of thrombin for its substrates. We analyzed polymerization, following changes in turbidity, and found that polymerization of A'beta fibrinogen was similar to that of normal fibrinogen. We analyzed clot structure by scanning electron microscopy and found that clots from A'beta fibrinogen were similar to clots from normal fibrinogen. We conclude that premature release of the fibrinopeptide from the N terminus of the beta chain does not affect polymerization of fibrinogen.  相似文献   

9.
Fibrinogen New York 1 (NY-1) was identified in a family with a thrombotic tendency. Studies on fibrinogen NY-1 and the fibrinogen from her brother, designated NY-1a, showed that both have abnormal thrombin-nonclottable fibrinogen (50% of the total fibrinogen in NY-1 and 35-40% in NY-1a) and that the trait is heterozygous and autosomal codominant. The abnormal fibrinogen polymerizes in the presence of calcium and can be further cross-linked by Factor XIIIa. The release rates of fibrinopeptides A and B by thrombin from both (NY-1 and NY-1a) were slower than those from normal fibrinogen. Two mol of fibrinopeptide A but only 0.6-1.0 (NY-1) or 1.0-1.3 (NY-1a) mol of fibrinopeptide B were released per mol of fibrinogen. Additionally, only 1.0 (NY-1) or 1.3 (NY-1a) mol of the B beta(1-42) peptide were released by plasmin/mol. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis of the reduced fibrinogen revealed two protein bands in the B beta-chain region (Mr = 54,000 as compared with 57,300 for the normal). When NY-1a fibrinogen was treated with CNBr, two sizes of the NH2-terminal disulfide knot were obtained (Mr = 59,000 and 49,000). The Mr = 49,000 component is consistent with an abnormal NH2-terminal disulfide knot with two defective NH2-terminal B beta-chains. Amino acid sequence analyses demonstrated that the abnormal B beta-chain is the result of a deletion in the sequence from residues 9 to 72. This deletion corresponds exactly to exon 2 of the gene. Since this family has a thrombotic tendency, the defect in the fibrinogen may be important in the pathogenesis of thrombosis in this family.  相似文献   

10.
Fibrin II induces endothelial cell capillary tube formation   总被引:11,自引:0,他引:11       下载免费PDF全文
We studied the formation of capillary tubes by endothelial cells which were sandwiched between two fibrin gels under serum-free conditions. After formation of the overlying fibrin gel, the endothelial cell monolayer rearranged into an extensive net of capillary tubes. Tube formation was apparent at 5 h and was fully developed by 24 h. The capillary tubes were vacuolated, and both intracellular and intercellular lumina were present. Maximal tube formation was observed with fibrin II (which lacks both fibrinopeptide A and B), minimal tube formation with fibrin I (which lacks only fibrinopeptide A), and complete absence of tube formation with fibrin 325 (which lacks the NH2- terminal beta 15-42 sequence, in addition to fibrinopeptides A and B). The inability of fibrin 325 to stimulate capillary tube formation supports the idea that beta 15-42 plays an important role in this process, and its importance was confirmed by the finding that exogenous soluble beta 15-42 inhibited fibrin II-induced capillary tube formation. This effect was specific for fibrin, since beta 15-42 did not inhibit tube formation by endothelial cells sandwiched between collagen gels. The interaction of the apical surface of the endothelial cell with the overlying fibrin II gel, as opposed to the underlying fibrin gel upon which the cells were seeded, was necessary for capillary tube formation. These studies suggest that the beta 15-42 sequence of fibrin interacts with a component of the apical cell surface and that this interaction plays a fundamental role in the induction of endothelial capillary tube formation.  相似文献   

11.
The formation of a fibrin clot occurs through binding of putative complementary sites, called fibrin polymerization sites, located in the NH2- and COOH-terminal domains of fibrin monomer molecules. In this study, we have investigated the structure of the NH2-terminal fibrin polymerization site by using fibrinogen-derived peptides and fragments. Fibrinogen was digested with Crotalus atrox protease III, to two major molecular species: a Mr 325,000 derivative (Fg325) and a peptide of Mr 5000. The peptide and its thrombin-cleavage product were purified by ion-exchange and reverse-phase HPLC; the authenticity of the B beta 1-42 and beta 15-42 peptides, respectively, was confirmed by amino acid sequencing. Since Fg325 had decreased thrombin coagulability, we addressed the question of whether the peptide B beta 1-42 contained a fibrin polymerization site. In order to identify and map the site, the peptides B beta 1-42 and beta 15-42 were tested for their ability to inhibit fibrin monomer polymerization. In addition the following peptides prepared by chemical synthesis were also tested: beta 15-18, beta 15-26, beta 24-42, beta 40-54, beta 50-55, and alpha 17-19-Pro. While B beta 1-42 had no inhibitory activity, the peptide devoid of fibrinopeptide B, beta 15-42, was a strong inhibitor. The peptides beta 15-18, beta 15-26, and beta 15-42 decreased the rate of fibrin polymerization by 50% at a molar excess of the peptide to fibrin monomer of 500, 430, and 50, respectively. The peptides beta 24-42, beta 40-54, and beta 50-55 were inactive.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
Thrombomodulin acts as a linear competitive inhibitor of thrombin with respect to the substrate fibrinogen. In the present study the effect of thrombomodulin on the activity of thrombin with fragments of the A alpha and B beta chain of fibrinogen has been examined. The cleavage of fibrinopeptide A from the N-terminal disulphide knot, fragment 1-44 and fragment 1-51 of the A alpha chain was inhibited by thrombomodulin. The average value for the inhibition constant obtained with these substrates was 0.83 +/- 0.09 nM, which was in good agreement with the values obtained previously for the inhibition of thrombin by thrombomodulin with native fibrinogen as the substrate [Hofsteenge, J., Taguchi, H. & Stone, S. R. (1986) Biochem. J. 237, 243-251]. In contrast, the cleavage of fibrinopeptide A from fragment 1-23 and fragment 1-29 of the A alpha chain was not affected by thrombomodulin. Although the cleavage of the B beta chain in the intact fibrinogen molecule was inhibited by thrombomodulin [Hofsteenge, J., Taguchi, H. & Stone, S. R. (1986) Biochem. J. 237, 243-251], the release of fibrinopeptide B from the N-terminal disulphide knot and the N-terminal 118-residue fragment of the B beta chain was not inhibited by thrombomodulin. In addition, we determined the second-order rate constants of cleavage of these substrates using human thrombin. Fragments of the A alpha chain whose cleavage was inhibited by thrombomodulin were found to have values for kcat/Km that were within one order of magnitude of that for the native fibrinogen, whereas those for A alpha chain fragments whose cleavage was not inhibited by thrombomodulin were found to be more than two orders of magnitudes lower. From these results we conclude that only a relatively small portion of the A alpha chain of the fibrinogen molecule is responsible for the specific binding to thrombin that is affected by thrombomodulin. Moreover, residues 30-44 of the A alpha chain play an important role in this thrombin-fibrinogen interaction.  相似文献   

13.
The conformations of the B beta chain of the intact fibrinogen molecule and of various fragments of the B beta chain of fibrinogen that contain the region that is hydrolyzed by thrombin have been compared by an immunochemical method [Sachs, D. H., Schechter, A. N., Eastlake, A., & Anfinsen, C. B. (1972) Proc. Natl. Acad. Sci. U.S.A. 69, 3790]. Anti-fibrinogen antibodies were induced in rabbits by immunization with native bovine fibrinogen. An antibody population specific for the native antigenic determinant within the B beta fragment 20-28 was isolated by immunoadsorption. This preparation was to determine the value of Kconf, the equilibrium constant for the interconversion of the nonnative and native conformations of this determinant. Values of Kconf were measured for this determinant within native fibrinogen, the disulfide knot (DSK), CNBrB beta, B beta fragment 16-28, B beta fragment 20-28, and fibrinopeptide B (FpB). 125I-Labeled fibrinogen (125I-F) was used in the determination of Kconf by measuring the competition between 125I-F and the fibrinogen derivatives under study for binding to the purified antibody. For the antigenic region in F, the DSK, and CNBrB beta, the values of Kconf at 4 degrees C were infinity, (5.9 +/- 3.5) X 10(-3), and (1.2 +/- 0.7) X 10(-3), respectively. The values of Kconf for B beta fragment 16-28, B beta fragment 20-28, and FpB at 4 degrees C were less than (6.0 +/- 3.9) X 10(-7).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
Human fibrinogen exposed to protease III from Crotalus atrox venom is cleaved near the NH2 terminus of the B beta chain yielding a species of Mr 325,000 (Fg325) with impaired thrombin clottability. The derivative was compared with intact fibrinogen in a number of ways to determine whether the functional defect resulted from a conformational change or from the loss of a polymerization site. NH2-terminal amino acid sequencing of isolated A alpha, B beta, and gamma chains showed that Fg325 contained intact A alpha and gamma chains, but differed from fibrinogen by the absence of the first 42 residues of the B beta chain. Fibrinopeptide A was present and was cleaved at the same rate in both fibrinogen and Fg325. The rate and extent of A alpha and gamma cross-linking by factor XIIIa was also indistinguishable. In contrast, the thrombin-catalyzed coagulation of Fg325 was 46% less in extent and 180-fold slower than observed for intact fibrinogen. A conformational comparison of Fg325 and fibrinogen was made using immunochemical and spectroscopic approaches. Antisera specific for different regions of the fibrinogen molecule were used to characterize the epitopes in Fg325. The only significant differences were found in the NH2-terminal region of the B beta chain, probed with antiserum to B beta 1-118. The conformational similarity of Fg325 and fibrinogen was confirmed by the identity of both near and far UV CD spectra of the two proteins. Structural, functional, and immunochemical results imply that cleavage of 42 NH2-terminal residues from the B beta chain is not accompanied by a measurable conformational change. The residues of this B beta chain segment, which are evidently located on the surface of the molecule, in conjunction with the NH2-terminal part of the A alpha chain appear to play an important role in the expression of a fibrin polymerization site.  相似文献   

15.
Family members heterozygous for the congenitally abnormal fibrinogen designated fibrinogen Manchester, A alpha 16Arg----His, have previously been shown by h.p.l.c. and amino acid analysis to release a variant fibrinopeptide, [His16]fibrinopeptide A, from plasma fibrinogen after the addition of thrombin. The present study was designed to determine if the same abnormal phenotype was also present in the intraplatelet fibrinogen pool. Fresh platelets were washed in buffers containing EDTA until it could be shown that all washable plasma fibrinogen was removed. Normal platelets were then lysed by freezing and thawing to release their intracellular proteins, which were then treated with thrombin. The fibrinopeptides, cleaved from the intraplatelet fibrinogen, could be detected by an optimized h.p.l.c. technique. Quantification of the intraplatelet fibrinogen gave a result (means +/- S.D., n = 5) of 110 +/- 30 and 90 +/- 30 micrograms/10(9) platelets, when determined by h.p.l.c. quantification of fibrinopeptide B content and fibrinogen fragment E radioimmunoassay respectively. Examination of fibrinopeptides released from the platelet fibrinogen from the family with fibrinogen Manchester with the same techniques showed elution peaks in the same positions as both [His16]fibrinopeptide A and normal fibrinopeptide A. The identity of these peaks was further substantiated by analysis of the h.p.l.c. peaks by using specific radioimmunoassay to fibrinopeptide A. Our results therefore demonstrate that platelet fibrinogen expresses the heterozygous A alpha 16His phenotype. This supports the view that the A alpha chains of platelet and plasma fibrinogen are produced from a single genetic locus.  相似文献   

16.
Previous studies have indicated that rabbit antisera R2 and R33 to human fibrinopeptide A differ markedly in terms of cross-reactivity with fibrinogen and fibrinopeptide A-containing fragments of the fibrinogen molecule. Antiserum specificity was characterized by comparison of inhibition of binding to radiolabeled tyrosyl fibrinopeptide A produced by synthetic fragments and enzymatic digests of the fibrinopeptide A molecule vs. the complete fibrinopeptide sequence (Aalpha 1-16). Synthetic COOH-terminal homologues through the dodecapeptide (Aalpha 5-16) exhibited less than 16% immunoreactivity with R33 antiserum, which cross-reacts extensively with fibrinogen and fibrinopeptide A-containing fibrinogen fragments. In contrast, the synthetic COOH-terminal decapeptide (Aalpha 7-16) gave 100% immunoreactivity with R2 antiserum, which cross-reacts minimally with fibrinogen and fibrinopeptide A-containing fibrinogen fragments. Synthetic homologues smaller than Aalpha 7-16, such as Aalpha9-16 and Aalpha 7-11, reacted only minimally with R2 antiserum. Carboxypeptidase B digests of fibrinopeptide A retained less than 25% of their initial immunoreactivity with R2 antiserum. It is concluded that the antigenic determinants of R2 immunoreactivity reside entirely within the COOH-terminal ten-residue sequence of fibrinopeptide A, and that Phe-8, Asp-7, and Arg-16 contribute significantly to R2 immunoreactivity. The R2 antigenic determinants appear to be significantly less accessible to reaction with antibody than the R33 determinants when the fibrinopeptide is attached to its parent alpha chain (Canfield et al., 1976). A possible mechanism for the sequestration is discussed.  相似文献   

17.
The cyanogen bromide fragment, N-DSK, containing the NH2-terminal portions of the three chains of fibrinogen, was found to exist in dimeric and polymeric forms. These different forms gave rise to identical chain fragments on reduction and alkylation. The B beta chain of N-DSK from fibrinogen and the beta chain of N-DSK from fibrin were isolated and characterized. The B beta chain fragment has a blocked NH2-terminal residue, and fibrinopeptide B is released on digestion with thrombin. The beta chain fragment has glycine as NH2-terminal residue. The molecular weight of the B beta chain fragment is 12200 as determined by ultracentrifugal analysis. Gel electrophoresis in sodium dodecyl sulphate gave the molecular weights of 14000 and 13000 for the B beta chain and beta chain fragments, respectively. The NH2-terminal B beta chain fragment consists of 118 amino acid residues and the beta chain fragment of 104 residues. The amino acid sequence of beta chain fragment is identical to B beta chain fragment except for the fibrinopeptide B portion. The isolation of a B beta-related fragment (B beta +), with a molecular weight of 30000, is also reported. The presence of B beta + was explained on the basis of incomplete cleavage at the Met-118 residue during treatment with cyanogen bromide. Some functional aspects of the B beta chain fragment are discussed.  相似文献   

18.
Staphylococcus epidermidis is an important opportunistic pathogen and is a major cause of foreign body infections. We have characterized the ligand binding activity of SdrG, a fibrinogen-binding microbial surface component recognizing adhesive matrix molecules from S. epidermidis. Western ligand blot analysis showed that a recombinant form of the N-terminal A region of SdrG bound to the native Bbeta chain of fibrinogen (Fg) and to a recombinant form of the Bbeta chain expressed in Escherichia coli. By analyzing recombinant truncates and synthetic peptide mimetics of the Fg Bbeta chain, the binding site for SdrG was localized to residues 6-20 of this polypeptide. Recombinant SdrG bound to a synthetic 25-amino acid peptide (beta1-25) representing the N terminus of the Fg Bbeta chain with a KD of 1.4 x 10(-7) m as determined by fluorescence polarization experiments. This was similar to the apparent K(D) (0.9 x 10(-7) m) calculated from an enzyme-linked immunosorbent assay where SdrG bound immobilized Fg in a concentration-dependent manner. SdrG could recognize fibrinopeptide B (residues 1-14), but with a substantially lower affinity than that observed for SdrG binding to synthetic peptides beta1-25 and beta6-20. However, SdrG does not bind to thrombin-digested Fg. Thus, SdrG appears to target the thrombin cleavage site in the Fg Bbeta chain. In fact, SdrG was found to inhibit thrombin-induced fibrinogen clotting by interfering with fibrinopeptide B release.  相似文献   

19.
In 64 patients affected with acute leukaemia (51 patients with acute non-lymphatic leukaemia and 13 patients with acute lymphatic leukaemia) extensive investigations of blood coagulation were made during cytostatic therapy. The following conspicuous changes of haemostatasis could be observed in making the diagnosis: Lowered quick value and shortened PTT, increased fibrinogen, fibrinopeptide, A, alpha 1-antitrypsin and alpha 2-macroglobulin, diminished plasminogen and plasma fibrininectin. According to TAD (VP) protocol the induction therapy leads to hypercoagulability which can be recognized by an increase of fibrinopeptide A, coagulating factors and shortening of PTT. During the therapy with L-asparaginasis procoagulatoric as well as thromboprotective coagulating proteins are diminished. A dense laboratory control enables those disturbances of haemotasis caused by disease or therapy to be separated and contributes to preventing complications during the cytostatic induction therapy.  相似文献   

20.
Factor VII Activating Protease (FSAP) is a plasma protease affecting both coagulation and fibrinolysis. Although a role in hemostasis is still unclear, the identification of additional physiologic substrates will help to elucidate its role in this context. FSAP has been reported to cleave fibrinogen, but the functional consequences of this are not known. We have therefore undertaken this study to determine the implications of this cleavage for fibrin-clot formation and its lysis. Treatment of human fibrinogen with FSAP released an N-terminal peptide from the Bβ chain (Bβ1-53) and subsequently the fibrinopeptide B; within the Aα chain a partial truncation of the αC-region by multiple cleavages was seen. The truncated fibrinogen showed a delayed thrombin-catalyzed polymerization and formed fibrin clots of reduced turbidity, indicative of thinner fibrin fibers. Confocal laser scanning and scanning electron microscopy of these clots revealed a less coarse fibrin network with thinner fibers and a smaller pore size. A lower pore size was also seen in permeability studies. Unexpectedly, FSAP-treated fibrinogen or plasma exhibited a significantly faster tPA-driven lysis, which correlated exclusively with cleavage of fibrinogen and not with activation of plasminogen activators. Similar observations were also made in plasma after activation of endogenous zymogen FSAP, but not in plasma of carrier of the rare Marburg I single nucleotide polymorphism. In conclusion, altering fibrin clot properties by fibrinogenolysis is a novel function of FSAP in the vasculature, which facilitates clot lysis and may in vivo contribute to reduced fibrin deposition during thrombosis.  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