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 562 毫秒
1.
The N-terminal portions of the Aα and Bβ chains of bovine fibrinogen (CNBr Aα and Bβ), each of which contains an ArgGly bond that is hydrolyzed by thrombin, have been isolated by cyanogen bromide cleavage of fibrinogen and column chromatography of the resulting material. These peptides were digested with thrombin, releasing fibrinopeptide A and GlyProArg from CNBr Aα, and fibrinopeptide B from CNBr Bβ. The C-terminal peptides produced by digestion with thrombin (CNBr α and CNBr β) were purified, and the amino acid sequences of portions of these peptides (30 residues from the N-terminus of CNBr α and 32 residues from the N-terminus of CNBr β) were determined with an automatic sequenator using the Edman degradation.  相似文献   

2.
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   

3.
The interaction of the following human fibrinogen-like peptides with bovine thrombin was studied by use of one- and two-dimensional NMR techniques in aqueous solution: Ala(1)-Asp-Ser-Gly-Glu-Gly-Asp-Phe(8)-Leu-Ala-Glu-Gly-Gly-Gly-Val-Arg(16 )- Gly(17)-Pro-Arg(19)-Val(20)-Val-Glu-Arg (F10), residues 1-16 of F10 (fibrinopeptide A), residues 17-23 of F10 (F12), residues 1-20 of F10 (F13), residues 6-20 of F10 with Arg(16) replaced by a Gly residue (F14), and residues 6-19 of F10 with Arg(16) replaced by a Leu residue (F15). At pH 5.3 and 25 degrees C, the Arg(16)-Gly(17) peptide bonds of both peptides F10 and F13 were cleaved instantaneously in the presence of 0.6 mM thrombin, whereas the cleavage of the Arg(19)-Val(20) peptide bonds in peptides F12, F13, and F14 took over 1 h for completion. On the basis of observations of line broadening, fibrinopeptide A was found to bind to thrombin. While resonances from residues Ala(1)-Glu(5) were little affected, binding of fibrinopeptide A to thrombin caused significant line broadening of NH and side-chain proton resonances within residues Asp(7)-Arg(16). There is a chain reversal within residues Asp(7)-Arg(16) such that Phe(8) is brought close to the Arg(16)-Gly(17) peptide bond in the thrombin-peptide complex, as indicated by transferred NOEs between the aromatic ring protons of Phe(8) and the C alpha H protons of Gly(14) and the C gamma H protons of Val(15). A similar chain reversal was obtained in the isolated peptide F10 at a subzero temperature of -8 degrees C. The titration behavior of Asp(7) in peptide F13 does not deviate from that of the reference peptide, N-acetyl-Asp-NHMe at both 25 and -8 degrees C, indicating that no strong interaction exists between Asp(7) and Arg(16) or Arg(19). Peptides with Arg(16) replaced by Gly and Leu, respectively, i.e., F14 and F15, were also found to bind to thrombin but with a different conformation, as indicated by the absence of the long-range NOEs observed with fibrinopeptide A. Residues Asp(7)-Arg(16) constitute an essential structural element in the interaction of thrombin with fibrinogen.  相似文献   

4.
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   

5.
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   

6.
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   

7.
To examine the role of protein catabolism in the formation of antigenic peptide fragments, human fibrinopeptide-immune guinea pig T cells were stimulated with the large native molecule, human fibrinogen. Two different systems were tested. In the first, we determined responses by human fibrinopeptide B (hFPB)-immune T cells, to which strain (St.) 2 guinea pigs are responders and St. 13 are nonresponders, and by human fibrinopeptide A (hFPA)-immune T cells to which St. 13 are responders and St. 2 are nonresponders. Of interest in this comparison is that both hFPA and hFPB are amino terminal peptides on the A and B chain of fibrinogen, respectively, and are readily cleaved by thrombin during fibrin formation and by other trypsin-like enzymes, leaving a carboxyl terminal Arg. Thus, if fibrinogen catabolism occurred, both antigenic peptides should be equally represented for availability in T cell responses. It was found that hFPB-immune St. 2 T cells responded to fibrinogen, but no response was observed with hPFA-immune St. 13 T cells cultured with fibrinogen. To rule out that there was a general catabolic defect in St. 13 antigen-presenting cells, fibrinogen was presented by (2 X 13)F1 macrophages to fibrinopeptide-immune parental T cells. Again it was found that F1 macrophages could present fibrinogen to hFPB-immune T cells but failed to present hFPA. In another comparison, responses with fibrinogen were also determined with des-ARg-hFPB, which lacks the carboxyl terminal Arg of hFPB, to which St. 13 are responders and St. 2 are nonresponders. The advantage of this comparison is that both antigenic determinants are contained within the same small peptide. St. 13 des-Arg-hFPB-immune T cells failed to respond in vitro by culture with human fibrinogen, suggesting that these antigenic determinants are not produced from larger peptides or proteins containing those determinants. To rule out the possibility that this was only an in vitro phenomenon, guinea pigs were immunized with the larger protein, the B chain of fibrinogen, and the immune T cells were examined for responses to fibrinopeptides derived from the B chain. Immune St. 2 T cells responded to hFPB but not to des-Arg-hFPB, whereas St. 13 T cells remained unresponsive with both peptides. These results indicate that proteolysis of larger proteins to form small antigenic peptides is not a random event and that not all potential antigenic determinants contained in a protein are produced during antigen processing.  相似文献   

8.
R Procyk  B Blomb?ck 《Biochemistry》1990,29(6):1501-1507
Fibrinogen contains 29 disulfide bonds. Limited reduction in buffers containing calcium led to cleavage of three of them: the two A alpha 442Cys-A alpha 472Cys intrapeptide disulfide bonds and the symmetrical A alpha 28Cys-A alpha 28Cys bond. The limited reduction did not affect clotting by thrombin. However, a prolongation of the thrombin clotting time occurred when the limited reduction took place in the absence of calcium. The bonds reduced under this condition included the three already mentioned and also the two gamma 326Cys-gamma 339Cys intrapeptide disulfide bonds located in the C-terminal ends of the gamma-chain. N-Terminal analysis of thrombin-treated samples showed that thrombin cleavage occurred at the normal A alpha 16-A alpha 17 site in fibrinogen that was partially reduced in the presence of calcium. By contrast, thrombin cleaved at the A alpha 19-A alpha 20 site in fibrinogen that was partially reduced in the absence of calcium, rendering the protein unclottable by removing the A alpha 17Gly-18Pro-19Arg peptide. The loss of thrombin clottability may have also come from gamma 326Cys-gamma 339Cys disulfide bond reduction since the structure supported by this bond may be important for the function of the C-terminal polymerization site. In samples of the partially reduced fibrinogen lacking the A alpha 17-19 residues, gel formation occurred through an oligomerization mechanism catalyzed by factor XIII.  相似文献   

9.
Z Zheng  R W Ashton  F Ni  H A Scheraga 《Biochemistry》1992,31(18):4426-4431
Fibrinogen Lille, a congenital dysfibrinogenemia, has been reported to arise from a mutation from Asp to Asn at position 7 of the A alpha chain of human fibrinogen, thereby reducing the thrombin-catalyzed rate of hydrolysis of the Arg(16)-Gly(17) peptide bond of this chain. Synthetic peptides of relevant portions of the wild-type and mutant A alpha chains were prepared, and the thrombin-catalyzed rates of hydrolysis of their Arg(16)-Gly(17) peptide bonds were determined. In addition, transferred NOE measurements were made to deduce their conformations, when complexed to bovine thrombin. The kinetics data showed little difference in the hydrolysis rates between the wild-type and mutant peptides, and the NMR data indicate no difference in the bound conformation of these two peptides. Therefore, electrostatic (or salt-bridge) interactions between Asp(7) and thrombin do not influence the bound conformations of these peptides. Asp(7) may interact with a remote residue of fibrinogen, not present in these synthetic peptides, or there may be additional mutations beyond A alpha (1-20) which have not been detected in fibrinogen Lille. Alternatively, when thrombin binds to fibrinogen at its secondary binding site, its primary (active) site may display different reactivities toward wild-type fibrinogen and fibrinogen Lille.  相似文献   

10.
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   

11.
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   

12.
We used fluorescence correlation spectroscopy (FCS) to study the activation of fibrinogen by thrombin and the subsequent aggregation of fibrin monomers into fibrin polymers at a very low and at physiological fibrinogen concentrations. In the labeling procedure used the fibrinogen was randomly labeled and the label was bound to the fibrinopeptide A and/or to the part of fibrinogen which after activation takes part in fibrin formation. We measured a diffusion coefficient for fibrinogen of 2.48 x 10(-7) +/- 0.10 x 10(-7) cm2/s. After activation with thrombin both fibrinopeptide A and fibrin polymerization products could be demonstrated. From our findings we suggest a model for the formation of a three-dimensional network as two parallel processes, elongation and branching and that fibrin oligomers are not only intermediates in the polymerization process but also are substrates for branching.  相似文献   

13.
The structure of the ternary complex of human alpha-thrombin with a covalently bound analogue of fibrinopeptide A and a C-terminal hirudin peptide has been determined by X-ray diffraction methods at 0.25 nm resolution. Fibrinopeptide A folds in a compact manner, bringing together hydrophobic residues that slot into the apolar binding site of human alpha-thrombin. Fibrinogen residue Phe8 occupies the aryl-binding site of thrombin, adjacent to fibrinogen residues Leu9 and Val15 in the S2 subsite. The species diversity of fibrinopeptide A is analysed with respect to its conformation and its interaction with thrombin. The non-covalently attached peptide fragment hirudin(54-65) exhibits an identical conformation to that observed in the hirudin-thrombin complex. The occupancy of the secondary fibrinogen-recognition exosite by this peptide imposes restrictions on the manner of fibrinogen binding. The surface topology of the thrombin molecule indicates positions P1'-P3', differ from those of the canonical serine-proteinase inhibitors, suggesting a mechanical model for the switching of thrombin activity from fibrinogen cleavage to protein-C activation on thrombomodulin complex formation. The multiple interactions between thrombin and fibrinogen provide an explanation for the narrow specificity of thrombin. Structural grounds can be put forward for certain congenital clotting disorders.  相似文献   

14.
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   

15.
Mullin JL  Gorkun OV  Lord ST 《Biochemistry》2000,39(32):9843-9849
We analyzed the polymerization of BbetaA68T fibrinogen, the recombinant counterpart of fibrinogen Naples, a variant known to have decreased thrombin binding. When polymerized with equal thrombin concentrations, BbetaA68T fibrinogen had a longer lag time and lower rate of lateral aggregation, V(max), than normal recombinant fibrinogen, but a similar final turbidity. At thrombin concentrations that equalized the rates of fibrinopeptide A release, BbetaA68T fibrinogen polymerized with a lag time and V(max) similar to normal, but reached a significantly lower final turbidity. Similar results were produced when BbetaA68T was polymerized with Ancrod, which cleaves fibrinopeptide A at the same rate from either fibrinogen, and when BbetaA68T desA monomers were polymerized. The polymerization of desAB fibrin monomers, which circumvents fibrinopeptide release, was the same for both fibrinogens. We confirmed that turbidity was indicative of fiber thickness by scanning electron microscopy of fibrin clots. Here, we present the first experimental evidence of fibrin polymerization with a normal period of protofibril formation and rate of lateral aggregation, but with a significantly decreased extent of lateral aggregation. We conclude that the decreased lateral aggregation seen in BbetaA68T fibrinogen is due to an altered step in the enzymatic phase of its polymerization process. We propose that during normal polymerization a subtle conformational change in the E domain occurs, between the release of FpA and FpB, and that this change modulates the mechanism of lateral aggregation. Without this change, the lateral aggregation of BbetaA68T fibrinogen is impaired such that variant clots have thinner fibers than normal clots.  相似文献   

16.
In an abnormal fibrinogen with impaired fibrin monomer polymerization designed as fibrinogen Osaka II, we have identified substitution of Arg by Cys at position 275 of the gamma chain. This Cys is linked to a free cysteine molecule by a disulfide link as evidenced by fast atom bombardment mass spectrometry. This finding was supported by identification of a single cysteine released from isolated abnormal fragment D1 upon reduction. This unique cystine structure at the mutation site has not been reported heretofore in any abnormal protein including fibrinogen. The substitution may well perturb the structure required for fibrin monomer polymerization, specifically that assigned to the carboxyl-terminal D domain of fibrinogen. Indeed, isolated fragment D1 with the Cys substitution failed to inhibit thrombin-mediated clotting of normal fibrinogen and normal fibrin monomer polymerization, while normal fragment D1 inhibited them markedly. Our data seem to provide supporting evidence that the putative polymerization site(s) assigned to the D domain of fibrinogen may be structure-dependent, including the carboxyl-terminal segment of the gamma chain as well as a contiguous region that contains the gamma 275 residue.  相似文献   

17.
1. Improved methods for the purification of lamprey thrombin and fibrinogen are presented. 2. Lamprey thrombin releases two fibrinopeptides from lamprey fibrinogen during the transformation into fibrin. Bovine thrombin releases only one of these, a peptide referred to as fibrinopeptide B. The differences in the by-products of fibrin formation are reflected in the different N-terminal amino acid compositions of the two types of fibrin. 3. The fibrinopeptide that is not removed from the lamprey fibrinogen by bovine thrombin can subsequently be released by treatment of that fibrin with lamprey thrombin. 4. Under the conditions used, lamprey thrombin releases both fibrinopeptides at about the same rate. 5. The differences in interaction among these pairs of related proteins are extreme manifestations of the phenomenon loosely referred to as `species specificity'.  相似文献   

18.
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   

19.
One of the peptides released from lamprey fibrinogen during its transformation into fibrin has been found to contain covalently bound carbohydrate. The peptide, which also contains tyrosine O-sulfate, corresponds to the mammalian fibrinopeptide B and is the amino-terminal segment of the lamprey fibrinogen β-chain. As noted previously, this peptide is the only one released when lamprey fibrinogen is coltted by mammalian thrombin. Of the more than fifty sets of fibrinopeptides characterized from various species, this is the first one found to contain carbohydrate.  相似文献   

20.
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   

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