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 171 毫秒
1.
E Mihalyi 《Biochemistry》1988,27(3):976-982
Kinetic data on the release of fibrinopeptides A and B from bovine fibrinogen by human thrombin were obtained at high fibrinogen concentrations, within the 0.8-8.8% range (0.227 X 10(-4) to 2.60 X 10(-4) M), at 25 degrees C, pH 7.26, 0.30 ionic strength, and 10(-4) M free Ca2+ concentration. Release of fibrinopeptide A followed strictly first-order kinetics at all concentrations, in spite of the fact that the highest concentration was 26 times larger than the value of KM found in the literature. This behavior can be explained by inhibition of thrombin by the reaction products, with KI = KM. The equation describing the course of the reaction under these conditions can be rearranged into a linear relationship between 1/kobsd and substrate concentration. The slope of the line is equal to 1/kcat and the intercept to KM/kcat. The data points fell accurately on a straight line, and with the parameters of the latter, kcat and KM were calculated as (6.3 +/- 0.11) X 10(-10) M s-1 (unit of thrombin)-1 L-1 and (11.0 +/- 3.0) X 10(-6) M, respectively. These values agree well with those found in the literature. Release of fibrinopeptide B follows complex kinetics. Higgins et al. [Higgins, D. L., Lewis, S.D., & Shafer, J.A. (1983) J. Biol. Chem. 258, 9276-9282] suggested that it can be described as the result of two consecutive reactions, the first one being the release of fibrinopeptide A and the second one of fibrinopeptide B from those molecules that have already lost fibrinopeptide A in the previous step.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
The time dependence of the release of fibrinopeptides from fibrinogen was studied as a function of the concentration of fibrinogen, thrombin, and Gly-Pro-Arg-Pro, an inhibitor of fibrin polymerization. The release of fibrinopeptides during fibrin assembly was shown to be a highly ordered process. Rate constants for individual steps in the formation of fibrin were evaluated at pH 7.4, 37 degrees C, gamma/2 = 0.15. The initial event, thrombin-catalyzed proteolysis at Arg-A alpha 16 to release fibrinopeptide A (kcat/Km = 1.09 X 10(7) M-1s-1) was followed by association of the resulting fibrin I monomers. Association of fibrin I was found to be a reversible process with rate constants of 1 X 10(6) M-1s-1 and 0.064 s-1 for association and dissociation, respectively. Assuming random polymerization of fibrin I monomer, the equilibrium constant for fibrin I association (1.56 X 10(7) M-1) indicates that greater than 80% of the fibrin I protofibrils should contain more than 10 monomeric units at 37 degrees C, pH 7.4, when the fibrin I concentration is 1.0 mg/ml. Association of fibrin I monomers was shown to result in a 6.5-fold increase in the susceptibility of Arg-B beta 14 to thrombin-mediated proteolysis. The 6.5-fold increase in the observed specificity constant from 6.5 X 10(5) M-1s-1 to 4.2 X 10(6) M-1s-1 upon association of fibrin I monomers and the rate constant for fibrin association indicates that most of the fibrinopeptide B is released after association of fibrin I monomers. The interaction between a pair of polymerization sites in fibrin I dimer was found to be weaker than the interaction of fibrin I with Gly-Pro-Arg-Pro and weaker than the interaction of fibrin I with fibrinogen.  相似文献   

3.
Steady-state kinetic parameters were compared for the action of alpha- and gamma-thrombin on the physiologically important thrombin substrates fibrinogen and factor XIII at 37 degrees C, pH 7.4, and 0.14 M NaCl. gamma-Thrombin, an alpha-thrombin derivative proteolytically cleaved at R-B73 and K-B154, was observed to catalyze the release of fibrinopeptide A (FPA) from fibrinogen with a specificity constant (kcat/Km) of 5 X 10(3) M-1 s-1. This value was approximately 2400-fold lower than the specificity constant for the corresponding alpha-thrombin-catalyzed reaction. The low specificity constant was attributed to an increase in Km and a decrease in kcat for gamma-thrombin-catalyzed release of FPA from fibrinogen. Conversion of alpha-thrombin to gamma-thrombin also resulted in an approximately 800-fold reduction in the specificity constant for thrombin-catalyzed release of fibrinopeptide B (FPB) from fibrin I, as well as a loss in discriminatory power. Whereas alpha-thrombin preferentially released FPA from intact fibrinogen, gamma-thrombin released FPA and FPB from intact fibrinogen at similar rates. In contrast to the large difference in specificity constants observed for alpha- and gamma-thrombin catalysis with fibrin(ogen) as substrate, the specificity constant (2.6 X 10(4) M-1 s-1) observed for gamma-thrombin-catalyzed release of activation peptide from factor XIII was only 5-fold lower than the corresponding value for the alpha-thrombin-catalyzed reaction. Additionally, the promotion of factor XIII activation by fibrin characteristic of the alpha-thrombin-catalyzed reaction did not occur in the gamma-thrombin-catalyzed reaction.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
R A Henriksen  K G Mann 《Biochemistry》1988,27(26):9160-9165
A congenitally dysfunctional form of prothrombin, prothrombin Quick, was isolated from the plasma of an individual with less than 2% of normal prothrombin activity. Following activation of prothrombin Quick, two dysfunctional thrombins, thrombin Quick I and thrombin Quick II, were isolated. Functional characterization of thrombin Quick I indicated an increase in KM and a decrease in kcat, relative to thrombin, for release of fibrinopeptide A. Comparison of kcat/KM for thrombin Quick I to the value obtained for thrombin yielded a relative catalytic efficiency of 0.012 for thrombin Quick I [Henriksen, R. A., & Owen, W. G. (1987) J. Biol. Chem. 262, 4664-4669]. Lysyl endopeptidase digestor of reduced and S-carboxymethylated thrombin and thrombin Quick I has resulted in the identification of an altered peptide in this dysthrombin. Edman degradation of the isolated peptide has shown that the altered residue in this protein is Arg-382 which is replaced by Cys. This could result from a point mutation in the Arg codon, CGC, to yield TGC. Together, these results indicate that Arg-382 is a critical residue in determining the specificity of thrombin toward fibrinogen. Similar relative activities for thrombin Quick I in stimulating platelet aggregation, in the release of prostacyclin from human umbilical vein endothelium, and in the release of fibrinopeptide A suggest that these activities of thrombin share the same specificity determinants.  相似文献   

5.
Pre-steady-state and steady-state kinetics of the papain (EC 3.4.22.2)-catalyzed hydrolysis of N-alpha-carbobenzoxyglycine p-nitrophenyl ester (ZGlyONp) have been determined between pH 3.0 and 9.5 (I = 0.1 M) at 21 +/- 0.5 degrees C. The results are consistent with the minimum three-step mechanism involving the acyl X enzyme intermediate E X P: (Formula: see text). The formation of the E X S complex may be regarded as a rapid pseudoequilibrium process; the minimum values for k+1 are 8.0 microM-1 s-1 (pH less than or equal to 3.5) and 0.40 microM-1 s-1 (pH greater than 6.0), and that for k-1 is 600 s-1 (pH independent). The pH profile of k+2/Ks (= kcat/Km; Ks = k-1/k+1) reflects the ionization of two groups with pK' values of 4.5 +/- 0.1 and 8.80 +/- 0.15 in the free enzyme. The pH dependence of k+2 and k+3 (measured only at pH values below neutrality) implicates one ionizing group in the acylation and deacylation step with pK' values of 5.80 +/- 0.15 and 3.10 +/- 0.15, respectively. As expected from the pH dependences of k+2/Ks (= kcat/Km) and k+2, the value of Ks changes with pH from 7.5 X 10(1) microM (pH less than or equal to 3.5) to 1.5 X 10(3) microM (pH greater than 6.0). Values of k-2 and k-3 are close to zero over the whole pH range explored (3.0 to 9.5). The pH dependence of kinetic parameters indicates that at acid pH values (less than or equal to 3.5), the k+2 step is rate limiting in catalysis, whereas for pH values higher than 3.5, k+3 becomes rate limiting. The observed ionizations probably reflect the acid-base equilibria of residues involved in the catalytic diad of papain, His159-Cys25. Comparison with catalytic properties of ficins and bromelains suggests that the results reported here may be of general significance for cysteine proteinase catalyzed reactions.  相似文献   

6.
Fibrinogen, purified from a recently identified case of dysfibrinogenaemia, fibrinogen Sydney I, was shown by thrombin digestion, high-performance liquid chromatography (HPLC) and amino acid analysis to be a heterozygous case of an A alpha Arg-16----His substitution. Kinetic studies have been carried out on the thrombin-induced release of fibrinopeptide A (FPA), fibrinopeptide B (FPB) and the variant peptide [His16]FPA. When thrombin was added to fibrinogen Sydney I at a concentration of 0.2 U/ml release of FPA was rapid and there was a 79-fold reduced rate of release of [His16]FPA, but the rate of release of FPB was not appreciably reduced. In contrast, at lower thrombin concentrations the rate of FPB release was reduced in proportion to the rate of total FPA release, supporting the view that release of fibrinopeptides is a sequential process. The second-order kinetic constant kcat/Km for hydrolysis of the abnormal A alpha chain by thrombin was calculated from Lineweaver-Burk plots to be 16-30-fold less than that for the normal A alpha chain. Molecular modelling studies, using a refined model of the trypsin-pancreatic-trypsin-inhibitor complex have been used to suggest how the histidine at the P1 site can be accommodated within the enzyme hydrophobic active-site pocket.  相似文献   

7.
Promotion of thrombin-catalyzed activation of factor XIII by fibrinogen   总被引:10,自引:0,他引:10  
T J Janus  S D Lewis  L Lorand  J A Shafer 《Biochemistry》1983,22(26):6269-6272
High-performance liquid chromatography was used to analyze the kinetics of the thrombin-catalyzed release of the activation peptide from the factor XIII zymogen (fibrin-stabilizing factor). The specificity constant (kcat/Km) for this reaction, measured at factor XIII concentrations much below Km, was (0.13-0.16) X 10(6) M-1 s-1 at pH 7.4, mu = 0.15, and 37 degrees C. Separate estimates, obtained from the dependence of the initial rates of release of the activation peptide on the concentration of factor XIII, gave values of 10 (+/- 3) s-1 for kcat and 84 (+/- 30) microM for Km, in terms of ab protomers of the zymogen. The thrombin-mediated release of the activation peptide was dramatically enhanced in the presence of fibrinogen. Furthermore, the time course of release, in relation to that of fibrinopeptide A, suggested that some des-A-fibrinogen species (e.g., alpha 2B beta 2 gamma 2) may be the true activator for promoting the cleavage of the Arg-36 peptide bonds in the a subunits of factor XIII. This observation suggests that generation of factor XIIIa and its substrate (fibrin) is coordinated so that thrombin-mediated zymogen activation proceeds efficiently only after the process of clotting has been initiated by the removal of fibrinopeptide A from fibrinogen.  相似文献   

8.
Mihalyi E 《Biophysical chemistry》2004,112(2-3):131-140
Calcium binding curves of human and bovine fibrinogen were obtained by using a calcium sensitive electrode. The two were identical and showed 2 high, 2-3 medium and more than 15 low affinity sites. Differential scanning calorimetry at neutral pH demonstrated the presence of the D and E domains of fibrinogen; however, at pH 3.5 the D-domain was split into two. The presence of the subdomains was demonstrated also by digestion by pepsin at this pH. Combination of digestion of fibrinogen and of its fragments with different enzymes and temperatures identified up to 12 subdomains in the original molecule. Clotting of fibrinogen by thrombin at pH 7.0 was investigated also by differential scanning calorimetry. In the absence of Ca2+ clotting elicited a 40% increase in the enthalpy of thermal denaturation of the D domain of fibrinogen, but the position of the peak increased only by 0.4 degrees C. However, with clotting in the presence of 10(-3) M calcium the former increased by 70-75% and the latter by 11.0 degrees C, while these parameters of the E-domain remained unchanged. Changes of bound calcium during clotting were also measured with the calcium sensitive electrode. These had to be corrected, because the drop in free calcium was partly compensated by release of some calcium that was already bound to fibrinogen. Log of the half time of calcium uptake plotted against log thrombin concentration indicated a first order process with respect to thrombin concentration, moreover, the rate determined corresponded to that of the conformation change measured by calorimetry. The calcium uptake was correlated with release of the fibrinopeptides. Release of fibrinopeptide B follows parallel to binding of calcium and that of fibrinopeptide A is about fourfold faster. Polymerization and formation of thick bundles of fibrin is connected with release of fibrinopeptide A. Clotting with Ancrod, an enzyme that releases only fibrinopeptide A, showed only minimal binding of calcium. The polymerization inhibiting tetrapeptide Gly-Pro-Arg-Pro also depressed binding of calcium. These data suggest that a calcium-binding site must be in the proximity of the site of release of fibrinopeptide B and of a polymerization site.  相似文献   

9.
The dysthrombin, Thrombin Quick, is chromatographically separable into two components designated Thrombin Quick I and Thrombin Quick II. Thrombin Quick II lacks observable catalytic activity toward thrombin substrates. The steady-state kinetics of hydrolysis of benzoylarginine ethyl ester and Tos-Gly-Pro-Arg-p-nitroanilide by Thrombin Quick I are equivalent to those of thrombin. These results, in addition to binding studies with the active site titrant N2-(5-dimethylaminonaphthalene-1-sulfonyl)arginine N-(3-ethyl-1,5-pentanediyl)amide, indicate that binding interactions at the catalytic site of Thrombin Quick I are unaltered. Thrombin Quick I is inhibited by anti-thrombin III at the same rate as thrombin. Steady-state kinetic parameters for the release of fibrinopeptide A indicate defects in both kcat and Km for Thrombin Quick I with kcat/Km equal to 0.012 of the value for thrombin, corresponding to the relative fibrinogen clotting activity of 0.013. The results are interpreted as indicating a defect in Thrombin Quick I at a binding site, external to the catalytic site, which is essential for determining specificity toward fibrinogen. The defect in kcat may result secondarily from small perturbations in the steric relationship of the catalytic triad residues. The rate of hydrolysis by Thrombin Quick I of the protein substrates bovine prothrombin and bovine protein C (in the absence of cofactors) is about one-third of that observed for thrombin, indicating that hydrolysis of these substrates by thrombin involves different specificity determinants than does the hydrolysis of fibrinogen.  相似文献   

10.
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   

11.
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   

12.
The kinetic parameters for release of fibrinopeptide A (FPA) from human fibrinogen by thrombin are: Km = 2.3 X 10(-6)M and Vmax. = 1.1 X 10(-10)mol of FPA/s per unit of thrombin; for fibrin formation, Km is similar to that for FPA release, but, the conditions of the present study, Vmax. was approximately half of that for FPA release. The formation of fibrin polymer before the sol-gel transition was studied by gel-permeation chromatography combined with effluent analysis for fibrinogen antigen and residual FPA. Polymer formation in purified fibrinogen incubated with thrombin proceeded as a bimolecular association of exposed sites in a manner predicted by probability calculations and assuming random FPA cleavage. Each oligomer consisted of n molecules of fibrin monomer and two fibrinogen molecules, each of the latter lacking one FPA molecule, i.e. each oligomer, regardless of molecular size, retains two FPA molecules. The addition of 5 mM-CaCl2 to the reaction mixture changed the rate of polymer formation, so that dimer was no longer the prevalent oligomer; in the presence of Ca2+, the trimer was the oligomer in highest concentration. The polymers formed in the presence of calcium were similar in composition to those without, i.e. 2 mol of FPA/mol of oligomer. EDTA-treated plasma samples incubated for short periods of time, 30s or less, with thrombin ranging in concentration up to 1 N.I.H. unit/ml did not form clots during the 10-15 min period of observation until they were applied to the column, though a large proportion of the available FPA was cleaved (maximum 45%). The soluble polymers in plasma were mostly of the high-Mr variety (tetramer and greater); these high-Mr polymers contained less than 2 mol of FPA/mol of polymer, whereas dimer and trimer in plasma were similar to those in the purified systems, i.e. 2 mol of FPA/mol.  相似文献   

13.
An enzyme bearing thrombin-like specificity has been purified to homogeneity from the venom of Trimeresurus flavoviridis (the Habu snake). The enzyme is a monomer with a molecular weight of 23,500 as determined by analytical gel filtration and sodium dodecyl sulfate polyacrylamide gel electrophoresis. The protein contains approximately 210 amino acid residues and has a relatively high content of aspartic acid and glutamic acid. The isoelectric point was 4.8 and the extinction coefficient at 280 nm for a 1% solution was 11.5. The enzyme acted directly on fibrinogen to form a fibrin clot with 2.0 NIH units. Analysis by high performance liquid chromatography of enzyme-treated fibrinogen revealed the release of a peptide identical in composition to thrombin-induced fibrinopeptide A, but no peptide corresponding to fibrinopeptide B was detected. The enzyme showed esterase and amidase activities on synthetic substrates containing arginine. The enzyme exhibited higher activity toward tosyl-L-arginine methyl ester (TAME) but 6-times lower activity toward benzoyl-L-arginine p-nitroanilide when compared with bovin thrombin. The esterase activity was inhibited by diisopropylfluorophosphate and at a slower rate by phenylmethanesulfonyl fluoride, but was least affected by tosyl-L-lysine chloromethyl ketone, showing that the enzyme is a serine protease like thrombin. The enzyme showed a bell-shaped pH dependence of kcat/Km for hydrolysis of TAME, with a maximum around pH 8.5.  相似文献   

14.
Human fibrinogen and the polymerization of fibrin after activation by the enzymes, thrombin and Batroxobin, were studied by means of dynamic laser light scattering (DLS). The apparent diffusion constant, D, for fibrinogen was measured and has a value of (1.80 +/- 0.42) X 10(-7) cm2 X s-1. D was found to contain contributions from the translational diffusion constant (Dt) as well as from the rotational diffusion constant (Dr). A comparison between experimental and calculated values of Dr and Dt suggests that fibrinogen in the absence of added Ca2+ expresses a certain degree of flexibility, while it is straightened in the presence of added Ca2+. The time dependence of D showed periodic oscillations, while the average D values decreased with time. Thrombin and Batroxobin caused similar behaviour of D. The period length was related to the enzyme concentration, clotting time (Ct) and the rate of release of fibrinopeptide A (FPA). No periodic oscillations were observed in experiments where the enzyme was replaced by saline, or in experiments using a dysfunctional fibrinogen (fibrinogen Aarhus) which displayed slow rates of FPA-release and polymerization. We propose that the periodic oscillations in a system far from equilibrium may be explained by conformational changes occurring in the fibrinogen molecule during enzyme activation and polymerization.  相似文献   

15.
Unsulfated N alpha-acetyl-hirudin45-65 (MDL 27 589), which corresponds to the C-terminus of hirudin1-65, was synthesized by solid-phase methods. The synthetic peptide was able to inhibit fibrin formation and the release of fibrinopeptide A from fibrinogen by thrombin. The catalytic site of thrombin was not perturbed by the synthetic peptide as H-D-Phe-Pip-Arg-pNA hydrolysis (amidase activity) was not affected. The binding of synthetic peptide and thrombin was assessed by isolation of the complex on gel-filtration chromatography. A single binding site with a binding affinity (Ka) of approx. 1.0 X 10(5) M-1 was observed for thrombin-hirudin45-65 interaction. The data suggest that the C-terminal residues 45-65 of hirudin contain a binding domain which recognizes thrombin and yet does not bind to the catalytic site of the enzyme.  相似文献   

16.
The alpha-carbonic anhydrase gene from Helicobacter pylori strain 26695 has been cloned and sequenced. The full-length protein appears to be toxic to Escherichia coli, so we prepared a modified form of the gene lacking a part that presumably encodes a cleavable signal peptide. This truncated gene could be expressed in E. coli yielding an active enzyme comprising 229 amino acid residues. The amino acid sequence shows 36% identity with that of the enzyme from Neisseria gonorrhoeae and 28% with that of human carbonic anhydrase II. The H. pylori enzyme was purified by sulfonamide affinity chromatography and its circular dichroism spectrum and denaturation profile in guanidine hydrochloride have been measured. Kinetic parameters for CO2 hydration catalyzed by the H. pylori enzyme at pH 8.9 and 25 degrees C are kcat=2.4x10(5) s(-1), KM=17 mM and kcat/KM=1.4x10(7) M(-1) x s(-1). The pH dependence of kcat/KM fits with a simple titration curve with pK(a)=7.5. Thiocyanate yields an uncompetitive inhibition pattern at pH 9 indicating that the maximal rate of CO2 hydration is limited by proton transfer between a zinc-bound water molecule and the reaction medium in analogy to other forms of the enzyme. The 4-nitrophenyl acetate hydrolase activity of the H. pylori enzyme is quite low with an apparent catalytic second-order rate constant, k(enz), of 24 M(-1) x s(-1) at pH 8.8 and 25 degrees C. However, with 2-nitrophenyl acetate as substrate a k(enz) value of 665 M(-1) x s(-1) was obtained under similar conditions.  相似文献   

17.
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   

18.
Structure-function relationships in the 6 epidermal growth factor-like domains of human thrombomodulin (TME, residues 227-462) were studied by deletion mutagenesis. Purified and characterised proteins were used for kinetic studies. Deletion of EGF1, EGF2 and residues 310-332 in EGF3 had no effect on thrombin binding (Kd) or on kcat/KM for protein C activation by the thrombin-thrombomodulin complex. Deletion of the rest of EGF3 and the interdomain loop between EGF3 and EGF4 had no effect on Kd but decreased kcat/KM to 10% of TME. Deletion of residues 447-462 of EGF6 had no effect on kcat/KM but increased Kd for thrombin approximately 6-fold. Thus, the region 333-350 in EGF3-4 is critical for protein C activation by the thrombin-thrombomodulin complex and the region 447-462 in EGF6 is critical for thrombin binding.  相似文献   

19.
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   

20.
Inhibition of the enzymatic activity of thrombin by concanavalin A   总被引:1,自引:0,他引:1  
Concanavalin A, a carbohydrate lectin derived from the jack bean, prolongs the thrombin clotting time of human plasma or purified fibrinogen. Prolongation is due to delay in peptide release from fibrinogen. The rate of fibrin monomer polymerization is not affected. Hydrolysis of protamine sulfate by thrombin is inhibited by concanavalin A. All inhibitory effects are prevented by α-methyl-D-mannoside. Concanavalin A does not delay clotting of fibrinogen by reptilase (releases fibrinopeptide A only) or by Ancistrodon contortrix contortrix (releases fibrinopeptide B initially followed by a small amount of A). It is concluded that concanavalin A binds to a carbohydrate on the thrombin molecule thus inhibiting its enzymatic activity.  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