首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
G Marx 《Biopolymers》1987,26(6):911-920
During the course of studies with fibrin protofibrils, produced by adding hirudin to thrombin-activated fibrinogen prior to the onset of gelation, turbid clots were observed to be generated merely by adding Ca(II) or Zn(II) to protofibrils. The rate of gelation (CT) and turbidity of the “protofibrin” clots increases with cation levels in a concentration-dependent manner, with Zn(II) much more potent than Ca(II). For example, 50 μM Zn(II) generated a more turbid protofibrin clot than 0.5 mM Ca(II). In combination, levels of Zn(II) and Ca(II), which individually have no effect, induce protofibril gelation. The generation of protofibrin clots by Zn(II) is decreased at increasing ionic strength. Apparently, the underlying electrostatic forces that bind the monomers in fibrin and protofibrin gels are similar. SEM micrographs show that Ca(II)- or Zn(II)-induced protofibrin clots (600–1500Å thick) are essentially indistinguishable from those formed directly from fibrinogen and thrombin with divalent cation. The protofibrin fibers induced by the cations are thicker than the fibers formed directly from fibrinogen and thrombin in the absence of divalent cation. Branching appears brought about the the divalent cation-sensitive lateral association of different protofibril strands. These findings describe simple experimental methods for separately studying the early and late stages of fibrin gelation.  相似文献   

2.
G Marx 《Biopolymers》1988,27(5):763-774
The cation sensitivity of linear and lateral assembly processes of thrombin- and reptilase-activated fibrinogen was examined. Analytic ultracentrifugation shows that the linear assembly of fibrin oligomers (protofibrils) is neither cation dependent nor sensitive to chelating agents. Protofibrils generated with thrombin–hirudin gelate with either 1–2 mM Ca(II) or 15–100 μM Zn(II). By contrast, protofibril B, generated with reptilase–diisopropylfluorophosphonate, gelates only with Ca(II) but is insensitive to Zn(II). These results indicate that the release of fibrinopeptides A and B (FPA and FPB) expose two types of lateral binding sites that are sensitive to Ca(II) and Zn(II) respectively. Transmission electron (TEM) micrographs of negatively stained gels indicate that the linear packing of the monomers within the fibrin- and cation-induced protofibrin fibers is essentially identical. Scanning electron (SEM) micrographs show that the Ca(II)-induced protofibrin B gel is similar to fibrin. In all, it seems that branching and gelation derive from two types of cation-sensitive, lateral associative processes. Based on these findings, a new paradigm for fibrin coagulation is proposed.  相似文献   

3.
Zinc binding to fibrinogen and fibrin was studied by two techniques. Scatchard analysis of ultrafiltration eluates reveals that fibrinogen has multiple Zn(II)-binding sites, KD (fibrinogen) = 18 microM; n = 6. The zinc content of the "collapsed" fibrin gel supernatant was also determined by atomic absorption spectroscopy and analyzed by a Scatchard plot (KD (fibrin) = 8 microM, n = 6). In other experiments, Zn(II) did not displace 45Ca(II) from fibrin. It appears that the binding of zinc to fibrinogen or fibrin is distinct from that of calcium, and that the zinc-binding characteristics of fibrinogen and fibrin are not significantly affected by the transformation of one into the other.  相似文献   

4.
Fibronectin and fibrin gel structure   总被引:4,自引:0,他引:4  
Plasma fibronectin is covalently incorporated into alpha-chains of fibrin gels in the presence of Factor XIII activated by thrombin (FXIIIaT) but not by Factor XIII activated by the snake venom enzyme batroxobin (FXIIIaB). FXIIIaB catalyzes introduction of gamma-gamma cross-links in fibrin but cross-linked alpha-chains are not formed. In the presence of FXIIIaT, fibrin gels formed by batroxobin incorporated fibronectin and the alpha-chains are cross-linked indicating that FXIIIaB has a different substrate specificity from FXIIIaT. In the presence of FXIIIaT the incorporation of fibronectin approaches 1 mol/340 kDa unit weight of fibrin. Fibronectin when present in a fibrinogen thrombin mixture containing FXIII does not influence the clotting time of the system nor the release of fibrinopeptides. Incorporation of fibronectin is not appreciable before the gel point. This indicates that the polymerization and gelation of fibrinogen is essentially not perturbed by the presence of fibronectin and that fibrin in the gel matrix rather than the fibrin polymers formed prior to gel point is the preferred structure for fibronectin incorporation. Incorporation of fibronectin into fibrin gels during formation leads to an increase in turbidity and a small decrease in Ks (permeability coefficient). This suggests that the width of the strands in the gel increases as a result of fibronectin incorporation. Fibronectin is also incorporated into preformed gels having completely cross-linked gamma- and alpha-chains perhaps indicating that the sites in fibrin involved in fibronectin incorporation are different from those involved in fibrin cross-linking. FXIIIaT appeared to be adsorbed to fibrin gel matrix in the presence but not in the absence of calcium ions.  相似文献   

5.
We identified a new property of human albumin. It enhances formation of fine fibril (or leptofibril) structures during fibrin gelation, and by nephelometric and electron microscopic measurements, this property is independent of and synergistic with that of fibrinogen. We examined fibrin aggregation using physiologic temperatures and pH and albumin:fibrin concentration ratios below those at which the known accelerating effect on fibrin aggregation occurs. An albumin concentration dependent decrease in gel turbidity maxima was consistently demonstrable in buffers containing or lacking (2-5 mM) CaCl2. This decrease was shown to be induced by albumin preparations which had been exposed to 2 mM ethylene-diaminetetraacetate disodium salt (EDTA), dialyzed, and tested in EDTA-free buffer. A delay in the onset of aggregation was also shown in calcium-lacking buffers by use of either reaggregating fibrin or fibrinogen aggregated with low (0.01-0.05 unit/mL) thrombin concentrations. Rates of fibrin aggregation as well as those of fibrinopeptide release were not affected by albumin, and the decrease in gel absorbance was demonstrable when solubilized fibrin was reaggregated at all final fibrin concentrations (0.2-4 microM) examined. Computed from wavelength dependence turbidity measurements (1 microM fibrin, I = 0.20), albumin decreased the average mass:length ratio from 8.24 X 10(11) to 4.26 X 10(11) daltons/cm, or from that of an approximately six to a three protofibril-thick strand. It also decreased the mean fibril radius from 48.5 to 36.4 nm but had no effect on fibril density. Electron microscopic measurements of cross-sectional fibril widths, performed on sections of glutaraldehyde-fixed gels, disclosed differences between albumin-containing and control gels which were significant by chi 2 analysis (P greater than 0.001). Fibril groups of 7-20- and 21-40-nm width together comprised 77% of fibrils formed in the presence of albumin (n = 251) compared to 30% of controls (n = 309). Conversely, coarser fibrils of 41-60- and 61-97-nm width together comprised 23% of fibrils formed in the presence of albumin and 70% of controls. This albumin effect was demonstrable by use of different monomeric albumin preparations including defatted, undefatted (unexposed and exposed to 60 degrees C, 10 h), chromatographically [gel exclusion and (diethylaminoethyl)cellulose] pure, S-(carboxymethyl)albumin, and S-(N-ethylsuccinimidyl)albumin. Chromatographically isolated albumin oligomers lacked this property, suggesting that a specific site(s) on albumin was (were) required.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

6.
Concanavalin A dimer interacts with fibrinogen and soluble fibrin at pH 5.2 Analysis of the binding data shows that there are in both cases four binding sites per molecule and that the dissociation constant does not change by removal of fibrinopeptides A and B. Ultracentrifugal studies shows that no aggregates of fibrinogen or fibrin are formed through concanavalin A binding and that up to four molecules of concanavalin A dimer can be bind to one molecule of fibrinogen or fibrin. These results imply that the four carbohydrate chains in the molecule are accessible to concanavalin A dimer. There is a diminution in the coagulation of fibrinogen by thrombin at low relative lectin concentrations and an increase at high concentrations. However, the lectin always favours the aggregation of fibrin monomers and does not have any inhibitory effect on the release of fibrinopeptides. We conclude that the electric charge in the neighbourhood of the carbohydrate in both chains, Bβ and γ plays an important role in the attraction between monomeric fibrin and fibrinogen-monomeric fibrin. The different effect of concanavalin A on the coagulation, depending on the relative concentration of the lectin, would be the result of the screening of this electric charge favouring either the interaction of fibrinogen-monomeric fibrin or the polymerization of monomeric fibrin.  相似文献   

7.
Watson JW  Doolittle RF 《Biochemistry》2011,50(45):9923-9927
Synthetic peptides patterned on sequences that appear during thrombin proteolysis of fibrinogen are known to influence fibrin formation in very different ways. A-Knob sequences (GPR-) inhibit polymerization, but B-knob sequences (GHR-) can actually enhance the process. We now report that when such peptides are attached to albumin carriers, both knob conjugates inhibit fibrin formation. In contrast, the 2-aminoethylthiol-albumin conjugate control enhances the polymerization to the same degree as albumin. The peptide AHRPam, which is known to bind exclusively to the βC holes of fibrinogen/fibrin, nullifies the inhibitory effects of the GHRPYGGGCam-albumin conjugate on fibrin polymerization, indicating that the inhibition was exclusively due to interactions with βC holes. AHRPam was much less effective in countering inhibition by the GPRPGGGGCam-albumin conjugate, suggesting that the observed effects with this conjugate involve mainly the γC holes of fibrin/fibrinogen. This study demonstrates that peptides modeled on fibrin polymerization knobs tethered to albumin retain their capacity to interact with fibrinogen/fibrin and may prove useful as inhibitors of clotting in vivo.  相似文献   

8.
Native fibrin gel networks observed by 3D microscopy, permeation and turbidity   总被引:15,自引:0,他引:15  
Native fully hydrated fibrin gels formed at different fibrinogen and thrombin concentrations and at different ionic strengths were studied by confocal laser 3D microscopy, liquid permeation and turbidity. The gels were found to be composed of straight rod-like fiber elements that often came together at denser nodes. In gels formed at high fibrinogen concentrations, or with high amounts of thrombin, the spaces between the fibers decreased, indicating a decrease of gel porosity. The fiber strands were also shorter. Gel porosity decreased dramatically in gels formed at the high ionic strengths. Shorter fibers were observed and fiber swelling occurred at ionic strengths above 0.24. Quantitative parameters for gel porosity, fiber mass/length ratio and diameter were also derived by liquid permeation and turbidometric analyses of the gels. Permeation analysis showed that gel porosity (measured as Ks) decreased in gels formed at higher fibrin and thrombin concentrations in agreement with the porosity observed by microscopy. The turbidometric analysis showed good agreement with the permeation data for gels formed at various thrombin concentrations, but supported the permeation data more poorly in gels formed at different fibrinogen concentrations, especially above 2.5 mg/ml. Turbidometric analysis showed that the fiber mass/length ratio and diameter decreased in gels formed at ionic strength up to 0.24, as was seen in the permeation study. However, at higher ionic strengths swelling of the fibers was suggested from the gel turbidity data and this was also indicated by microscopy. These findings are discussed in relation to previous hydrodynamic and electron microscopic studies of fibrin gels.  相似文献   

9.
Holographic relaxation spectroscopy (HRS) has been used to study transport of benzospiropyran (SP), BSA labeled with azobenzene (BSA-ABITC), and IgG-κ labeled with fluorescein (IgG-FITC) through fibrin gels formed under various conditions. The structures of the gels were controlled by means of the concentrations of fibrinogen, thrombin, and Ca2+ present during assembly of the fibrin. The diffusion coefficient of free dye (SP) was found to be independent of the fibrinogen concentration. The diffusion rate of labeled BSA reflected the assembly conditions of the gel for fibrinogen concentrations above approximately 6 g/L. In particular, the diffusion coefficient was higher in gels formed in the presence of 5 mM Ca2+. The labeled IgG showed photoinduced aggregation, as previously reported, as well as photoinduced attachment to the gel network to produce a permanent diffraction grating. Thus IgG is not a probe in the classical sense, but provides a model for protein diffusion and interactions in gels. These studies indicate that HRS is well suited to the study of molecular transport in fibrin gels.  相似文献   

10.
Albumin Tokushima is a Japanese genetic variant of human serum albumin. Two homozygous and 6 heterozygous subjects with this variant were found in a family. Albumin Tokushima was purified from sera of the homozygous subjects. Its amino acid composition and amino-terminal sequence were determined and compared with those of a normal serum albumin. Albumin Tokushima with the amino-terminal sequence of Arg-Gly-Val-Phe-His-Arg-Asp-Ala-His-Lys-Ser-Glu-Val-Ala-His-Arg-Phe-Lys- Asp- Leu-Gly-Glu-Glu-Asn-Phe was found to be the same abnormal proalbumin as proalbumin Lille (Abdo, Y. et al. (1981) FEBS Lett. 131, 286-288). The isoelectric points of albumin Tokushima were pH 4.70 and 4.90 as compared with pH 5.05 and 5.25 of a normal serum albumin. Albumin Tokushima was converted to normal serum albumin by purified cathepsin B in vitro. Albumin Tokushima can bind Ni2+ at 4 degrees C but binds little at 37 degrees C.  相似文献   

11.
Summary This study deals with the role of the mechanical properties of matrices in in vitro angiogenesis. The ability of rigid fibrinogen matrices with fibrin gels to promote capillarylike structures was compared. The role of the mechanical properties of the fibrin gels was assessed by varying concentration of the fibrin gels. When the concentration of fibrin gels was decreased from 2 mg/ml to 0.5 mg/ml, the capillarylike network increased. On rigid fibrinogen matrices, capillarylike structures were not formed. The extent of the capillarylike network formed on fibrin gels having the lowest concentration depended on the number of cells seeded. The dynamic analysis of capillarylike network formation permitted a direct visualization of a progressive stretching of the 0.5 mg/ml fibrin gels. This stretching was not observed when fibrin concentration increases. This analysis shows that 10 h after seeding, a prearrangement of cells into ringlike structures was observed. These ringlike structures grew in size. Between 16 and 24 h after seeding, the capillarylike structures were formed at the junction of two ringlike structures. Analysis of the αvβ3 integrin localization demonstrates that cell adhesion to fibrinogen is mediated through the αvβ3 integrin localized into adhesion plaques. Conversely, cell adhesion to fibrin shows a diffuse and dot-contact distribution. We suggest that the balance of the stresses between the tractions exerted by the cells and the resistance of the fibrin gels triggers an angiogenic signal into the intracellular compartment. This signal could be associated with modification in the αvβ3 integrin distribution.  相似文献   

12.
A Shimizu  J D Ferry 《Biopolymers》1988,27(4):703-713
Human fibrinogen (concentration 8.4 mg/mL) was ligated (cross-linked) with factor XIIIa and dithiothreitol (DTT) at pH 8.5, ionic strength 0.45. With 7.5 μg/mL of factor XIIIa alone, there was almost no γ-γ ligation, but with 2 mM DTT added, oligomers appeared, and γ-γ and Aα-Aα ligation was nearly complete after 3 days. At 38 μg/mL of factor XIIIa, some γ-γ and Aα-Aα ligation occurred even without DTT. For fibrinogen concentrations of 4.0 and 8.4 mg/mL, 38 μ/mL factor XIIIa, 2.0 mM DTT, clot-like gels formed and the shear modulus of elasticity increased slowly over several days to a constant value. The final modulus was similar in magnitude to those of ligated clots of α-fibrin (clotted by thrombin) and α-fibrin (clotted by batroxobin) under the same conditions. However, the opacity was somewhat higher; whereas in fine fibrin clots there is minimal lateral association of the protofibrils, in fibrinogen gels at the same pH and ionic strength the protofibrils (which are presumably single chains of fibrinogen monomers joined end to end at their D domains) are evidently associated in bundles (although not to the degree seen in coarse fibrin clots). Creep and creep recovery measurements showed almost perfect elastic behavior, with essentially no creep under stress and complete recovery after removal of stress. The modulus was scarcely affected by introduction of lithium bromide by diffusion to a concentration of 0.6M, which in unligated fibrin clots causes substantial softening. Whereas in fine fibrin clots (both αβ-fibrin and α-fibrin) factor XIIIa causes only γ-γ ligation, addition of 2 mM DTT produced some α-α ligation in these also.  相似文献   

13.
L Orning  F A Fitzpatrick 《Biochemistry》1992,31(17):4218-4223
Albumins from several species activated the bifunctional, Zn2+ metalloenzyme amino-peptidase/leukotriene A4 hydrolase (EC 3.3.2.6). Bovine serum albumin, 1 mg/mL, increased hydrolysis of L-proline-p-nitroanilide and leucine-enkephalin by 12-fold and 7-fold, respectively. The apparent Km for L-proline-p-nitroanilide was inversely proportional to the albumin concentration from 0 to 1 mg/mL, declining from 9.4 to 0.7 mM without an appreciable change in apparent Vmax. These data imply a random activation process in which the enzyme-activator complex is catalytically dominant. Hill plots indicated a 1:1 stoichiometric relationship between albumin and enzyme. Secondary plots of slope versus the reciprocal of albumin concentration indicated that it binds to the enzyme with an affinity constant of 0.9 microM. The pH optimum of the nonactivated enzyme occurred at pH 8; the albumin-activated enzyme had an optimum near pH 7. Neither ultrafiltration nor dialysis of albumin altered its activating effect, but boiling abolished it. Albumin did not affect other cytosolic or microsomal leucine aminopeptidases, or gamma-glutamyltransferase. Albumin functions as a nonessential activator, since enzymatic activity was always detectable in its absence. Chloride ions, which activate other Zn2+ metalloenzymes, also activated leukotriene A4 hydrolase/aminopeptidase with an EC50 = 50 mM, increasing its initial velocity 2.2-fold in the absence of albumin. Zn2+ activated the enzyme, increasing its apparent Vmax but not its apparent Km, suggesting it replaced Zn2+ lost from the active site, especially at acidic pH. At concentrations greater than 30-50 microM, Zn2+ was inhibitory. Albumin mitigated the effect of chloride, but not the effect of Zn2+ or that of the competitive inhibitor, captopril.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
Vitronectin is an abundant plasma protein that regulates coagulation, fibrinolysis, complement activation, and cell adhesion. Recently, we demonstrated that plasma vitronectin inhibits fibrinolysis by mediating the interaction of type 1 plasminogen activator inhibitor with fibrin (Podor, T. J., Peterson, C. B., Lawrence, D. A., Stefansson, S., Shaughnessy, S. G., Foulon, D. M., Butcher, M., and Weitz, J. I. (2000) J. Biol. Chem. 275, 19788-19794). The current studies were undertaken to further examine the interactions between vitronectin and fibrin(ogen). Comparison of vitronectin levels in plasma with those in serum indicates that approximately 20% of plasma vitronectin is incorporated into the clot. When the time course of biotinylated-vitronectin incorporation into clots formed from (125)I-fibrinogen is monitored, vitronectin incorporation into the clot parallels that of fibrinogen in the absence or presence of activated factor XIII. Vitronectin binds specifically to fibrin matrices with an estimated K(d) of approximately 0.6 microm. Additional vitronectin subunits are assembled on fibrin-bound vitronectin multimers through self-association. Confocal microscopy of fibrin clots reveals the globular vitronectin aggregates anchored at intervals along the fibrin fibrils. This periodicity raised the possibility that vitronectin interacts with the gamma A/gamma' variant of fibrin(ogen) that represents about 10% of total fibrinogen. In support of this concept, the vitronectin which contaminates fibrinogen preparations co-purifies with the gamma A/gamma' fibrinogen fraction, and clots formed from gamma A/gamma' fibrinogen preferentially bind vitronectin. These studies reveal that vitronectin associates with fibrin during coagulation, and may thereby modulate hemostasis and inflammation.  相似文献   

15.
Peptides that bind to fibrin but not to fibrinogen or serum albumin were selected from phage display libraries as targeting moieties for thrombus molecular imaging probes. Three classes of cyclic peptides (cyclized via disulfide bond between two Cys) were identified with consensus sequences XArXCPY(G/D)LCArIX (Ar = aromatic, Tn6), X(2)CXYYGTCLX (Tn7), and NHGCYNSYGVPYCDYS (Tn10). These peptides bound to fibrin at ~2 sites with K(d) = 4.1 μM, 4.0 μM, and 8.7 μM, respectively, whereas binding to fibrinogen was at least 100-fold weaker. The peptides also bind to the fibrin degradation product DD(E) with similar affinity to that measured for fibrin. The Tn7 and Tn10 peptides bind to the same site on fibrin, while the Tn6 peptides bind to a unique site. Alanine scanning identified the N- and C-terminal ends of the Tn6 and Tn7 peptides as most tolerant to modification. Peptide conjugates with either fluorescein or diethylenetriaminepentaaceto gadolinium(III) (GdDTPA) at the N-terminus were prepared for potential imaging applications, and these retained fibrin binding affinity and specificity in plasma. Relaxivity and binding studies on the GdDTPA derivatives revealed that an N-terminal glycyl linker had a modest effect on fibrin affinity but resulted in lower fibrin-bound relaxivity.  相似文献   

16.
Siebenlist KR  Meh DA  Mosesson MW 《Biochemistry》2000,39(46):14171-14175
There are conflicting ideas regarding the location of the carboxyl-terminal regions of cross-linked gamma-chain dimers in double-stranded fibrin fibrils. Some investigators believe that the chains are always oriented longitudinally along each fibril strand and traverse the contacting ends of abutting fibrin D domains ("DD-long" cross-linking). Other investigations have indicated instead that the chains are situated transversely between adjacent D domains in opposing fibril strands (transverse cross-linking). To distinguish between these two possibilities, the gamma dimer composition of factor XIIIa-cross-linked fibrin/fibrinogen complexes that had been formed through noncovalent D/E interactions between fibrinogen D domains and fibrin E domains was examined. Two factor XIIIa-mediated cross-linking conditions were employed. In the first, fibrin/fibrinogen complexes were formed between (125)I-labeled fibrinogen 2 ("peak 2" fibrinogen), each heterodimeric molecule containing one gamma(A) and one larger gamma' chain, and nonlabeled fibrin 1 molecules ("peak 1" fibrin), each containing two gamma(A) chains. If DD-long cross-linking occurred, (125)I-labeled gamma(A)-gamma(A), gamma(A)-gamma', and gamma'-gamma'dimers in a 1:2:1 ratio would result. Transverse cross-linking would yield a 1:1 mixture of (125)I-labeled gamma(A)-gamma(A) and gamma(A)-gamma' dimers, without any gamma'-gamma' dimers. Autoradiographic analyses of reduced SDS-PAGE gels from protocol 1 revealed (125)I-labeled gamma(A)-gamma(A) and gamma(A)-gamma' dimers at a ratio of approximately 1:1. No labeled gamma'-gamma' dimers were detected. Protocol 2 used a converse mixture, (125)I-fibrin 2 and nonlabeled fibrinogen 1. DD-long cross-linking of this mixture would yield only nonradioactive gamma(A)-gamma(A) dimers, whereas transverse cross-linking would yield a 1:1 mixture of (125)I-labeled gamma(A)-gamma(A) and gamma(A)-gamma' dimers. Autoradiographic analyses of this mixture yielded (125)I-labeled gamma(A)-gamma(A) and gamma(A)-gamma' dimers in a 1:1 ratio. These findings provide no evidence that longitudinal (DD-long) gamma chain positioning occurs in cross-linked fibrin and indicate instead that most, if not all, gamma-chain positioning in an assembled fibrin polymer is transverse.  相似文献   

17.
alpha-Sarcin binds one Zn(II) cation per protein molecule, with a Kd value of 0.9 mM, determined by equilibrium dialysis experiments. Ca(II), Mg(II), and Mn(II) do not bind to alpha-sarcin. Cd(II) and Co(II) also behave as Zn(II). The binding produces local modifications on the protein conformation affecting the microenvironment of tryptophan residues. The three cations modify the fluorescence emission of the protein. The near-u.v. circular dichroism spectrum of the protein is also altered. The binding of Zn(II) and related cations does not modify the secondary structure of the protein. The ribonucleolytic activity of alpha-sarcin is inhibited upon Zn(II) binding, but no alteration of the ability of the protein to aggregate phospholipid vesicles has been observed.  相似文献   

18.
Localization of a fibrin polymerization site   总被引:6,自引:0,他引:6  
The formation of a fibrin clot is initiated after the proteolytic cleavage of fibrinogen by thrombin. The enzyme removes fibrinopeptides A and B and generates fibrin monomer which spontaneously polymerizes. Polymerization appears to occur though the interaction of complementary binding sites on the NH2-terminal and COOH-terminal (Fragment D) regions of the molecule. A peptide has been isolated from the gamma chain remnant of fibrinogen Fragment D1 which has the ability to bind to the NH2-terminal region of fibrinogen as well as to inhibit fibrin monomer polymerization. The peptide reduces the maximum rate and extent of the polymerization of thrombin or batroxobin fibrin monomer and increases the lag time. The D1 peptide does not interact with disulfide knot, fibrinogen, or Fragment D1, but it binds to thrombin-treated disulfide knot with a Kd of 1.45 X 10(-6) M at approximately two binding sites per molecule of disulfide knot. Fibrin monomer formed either by thrombin or batroxobin binds approximately two molecules of D1 peptide per molecule of fibrin monomer, indicating that the complementary site is revealed by the loss of fibrinopeptide A. The NH2-terminal sequence (Thr-Arg-Trp) and COOH-terminal sequence (Ala-Gly-Asp-Val) of the D1 peptide were determined. Therefore the gamma 373-410 region of fibrinogen contains a polymerization site which is complementary to the thrombin-activated site on the NH2-terminal region of fibrinogen.  相似文献   

19.
When fibrin polymerizes in a strong magnetic field, it can be highly oriented. The structural diffraction study of the oriented polymer becomes thus possible. The magnetic birefringence can also be used to study the development of the polymer Fibrinogen in solution is weakly oriented in high magnetic fields. In this work we present complementary results and discussion. The validity of the comparison of the orientation parameters of fibrinogen and fibrin with those of other orientable known biological structures is discussed. The orientation of fibrin formed from fibrin monomer solution is compared to that of fibrin formed by the action of thrombin on fibrinogen. The conditions to obtain highly oriented fibrin gels suitable for three dimensional structure studies are also briefly discussed.  相似文献   

20.
We studied the effects of contact of bovine pulmonary artery endothelial cell monolayers with fibrin on the endothelial barrier function. Fibrin formed by clotting purified fibrinogen (0.5 to 3.0 mg/ml) with alpha-thrombin (1 U/ml) was added to endothelial monolayers and permeability measurements were made after fibrin removal. Fibrin incubation for 3 hours resulted in 2- to 5-fold increases in transendothelial 125I-albumin permeability. Permeability returned to baseline value within 3 hours after fibrin removal. Direct contact with fibrin was necessary for the response, since fibrin separated from the endothelium did not increase permeability. Contact with agarose (2 mg/ml) or fibrinogen (0.5 to 3.0 mg/ml) also did not increase endothelial permeability. Transmission electron microscopic examination indicated normal appearance of interendothelial junctions at a time when albumin permeability was increased and no overt evidence of endothelial injury. Incubation of fibrin with endothelial monolayers at 4 degrees C prevented the increase in albumin permeability. We examined the possibility that increased albumin transcytosis was responsible for fibrin's effect using 14C-sucrose (Mr = 342D), a lipid insoluble tracer. Fibrin increased sucrose flux by 1.5-fold compared to 2- to 5-fold increases in albumin flux. The results indicate that fibrin contact with the endothelial cell increases endothelial permeability. The effect of fibrin may involve activation of temperature-sensitive bulk phase transcytosis of albumin.  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