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.
Previously, we reported that the glycosaminoglycan (GAG) hyaluronic acid (HA) specifically bound to the plasma protein fibrinogen [LeBoeuf, R. D., Raja, R. R., Fuller, G. M., & Weigel, P. H. (1986) J. Biol. Chem. 261, 12586]. The binding of other macromolecules to fibrinogen could influence the conversion of fibrinogen to fibrin. Therefore, we tested whether HA and other GAGs could alter the kinetics of fibrin polymer formation and the physical structure of the resulting gel. In this study, we present data showing that the GAGs HA and chondroitin sulfate (CS) affect fibrin formation in three specific ways: (i) they decreased the clotting time of fibrinogen 3-10-fold; (ii) both GAGs increase significantly the rate of fibrin polymer formation; and (iii) fibrin gels containing HA or CS had a final A450 that was greater than controls, indicating that these two glycosaminoglycans influence either the final size of fibrin fibrils or the extent of the lateral association between fibrils. These results demonstrate that the interactions of HA and CS with forming fibrin polymers can alter both the kinetics of formation and may produce structural changes in fibrin gels.  相似文献   

2.
G Marx  N Harari 《Biochemistry》1989,28(20):8242-8248
Albumin modulation of fibrin and protofibrin coagulation parameters was studied. Cation-depleted, fatty acid free, human and bovine albumins decrease fibrin clot turbidity in a concentration-dependent manner. Albumin also inhibits the formation of protofibrin gels, induced by addition of 25 microM Zn(II) to protofibrils, though it does not bind to (proto)fibrin. In order to verify that competition for cations underlies the influence of albumin, fibrinogen was dialyzed against cation-depleted albumin. Elemental analysis indicates a redistribution of Zn(II) from the fibrinogen to the albumin compartment, and the resultant fibrin clots are less turbid. Apparently, cation-depleted albumin acts as a competitor for divalent cations. The ability of albumin to compete for available Zn(II) was also expressed in gels formed by pH-jump experiments, in which fibrin monomer, maintained soluble at pH 4.9, is induced to change phase by addition of NaOH to pH 7.4. While turbidimetric evidence indicates that individual fibrin fibers simply become thinner with albumin, scanning electron micrographs (SEM) reveal a more complex effect on ultrastructure. Though albumin does not bind to the gels, fibrin gels produced with albumin show major changes in fiber ultrastructure, particularly evident in gels formed in the presence of Zn(II). These structural modifications are discussed within the context of the "excluded volume" effect, in which "crowding" by albumin alters (proto)fibrin reactivity and ultrastructure.  相似文献   

3.
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   

4.
Light scattering from fibrin structures, obtained by exposure of fibrinogen to thrombin, Batroxobin (Reptilase) or coagulant fraction extracted from Contortrix venom at 20 and and 37°C, show in every case that rod-like intermediates are formed in the beginning of the aggregation process. The fibrils differ in the extent of branching and in lateral aggregation. Contortrix enzyme causes the highest branching density but the lowest lateral aggregation. Thrombin and Batroxobin give almost identical results. A change of temperature from 20 to 37°C yields an increase in branching density and lateral aggregation for the fibrin structures induced by the two snake venoms. With thrombin, however, the branching density decreases with the elevated temperature while the lateral aggregation strongly increases. Mostly opaque clots are obtained, with the exception of the clots induced by thrombin at 37°C, where a fine or traslucent gel is obtained. A very low extent of branching and translucent gels are also found with thrombin at 20°C and pH 7.3 but at pH 9.5 no correlation between a preferential cleavange of fibrinopeptide B and the lateral aggregation could be detected. The opacity is discussed as being the result of inhomogeneity in both branching and lateral aggregation. A quantitative analysis of the angular dependence of the scattered light indicates that non-activated human fibrinogen exists at least in the two conformations of a long rod, L = 95 ± 5 nm, and a short rod of 47.5 ± 5 nm, with mass fractions of ~ 70 and 30%, respectively. Only the long rod conformation of the monomer is built in the fibril. The model of a pure end-to-end aggregation is shown to be unlikely and the possibility of an overlapping of the monomeric rods over a region of ~ 8 nm is discussed.  相似文献   

5.
The anisotropic mechanical properties of magnetically aligned fibrin gels were measured by magnetic resonance elastography (MRE) and by a standard mechanical test: unconfined compression. Soft anisotropic biomaterials are notoriously difficult to characterize, especially in vivo. MRE is well-suited for efficient, non-invasive, and non-destructive assessment of shear modulus. Direction-dependent differences in shear modulus were found to be statistically significant for gels polymerized at magnetic fields of 11.7 and 4.7 T compared to control gels. Mechanical anisotropy was greater in the gels polymerized at the higher magnetic field. These observations were consistent with results from unconfined compression tests. Analysis of confocal microscopy images of gels showed measurable alignment of fibrils in gels polymerized at 11.7 T. This study provides direct, quantitative measurements of the anisotropy in mechanical properties that accompanies fibril alignment in fibrin gels.  相似文献   

6.
1. The effects of acid mucopolysaccharides and acid mucopolysaccharide-proteins on the size and rate of formation of fibril aggregates from collagen solutions in pH7.6 buffers were studied by turbidimetric and light-scattering methods. 2. Serum albumin, orosomucoid, methylated cellulose, chondroitin sulphate A and chondroitin sulphate C of molecular weight less than 20000, and hyaluronate of molecular weight less than 40000 did not influence rates of fibril formation. Chondroitin sulphate A, chondroitin sulphate C and hyaluronate of high molecular weight retarded the rate of fibril formation. This effect of high-molecular-weight chondroitin sulphate C decreased with increasing ionic strength. Heparin, though of low molecular weight (13000), was highly effective, as was also heparitin sulphate. The chondroitin sulphate-proteins of very high molecular weight were highly effective, despite the fact that for some preparations the component chondroitin sulphate chains had molecular weights much less than 20000. 3. Agents that had delayed fibril formation were also effective in producing an increase in degree of aggregation of fibrillar collagen, as indicated by dissymmetry changes observed in light-scattering experiments at low collagen concentrations. Methylated cellulose and heparin at 2.5mug./ml. were unusual in decreasing aggregation, but heparin at 0.25mug./ml. increased aggregation. Electron microscopy of gels showed fibrils and fibril aggregates with ;normal' collagen spacing and dimensions consistent with the light-scattering results. 4. The rates of electrical transport of agents and of solvent (electro-osmosis) through collagen gels indicated a contribution of molecular entanglement that increased with increase in molecular size of the agents. Electrostatic binding of heparin to collagen was noted. Binding to collagen during fibril formation was also found for heparitin sulphate and a chondroitin sulphate with extra sulphate groups. 5. Electrostatic binding of acid mucopolysaccharide-proteins to collagen may be an important factor in the organization and functioning of connective tissues at all stages of growth and development. Excluded-volume (molecular-entanglement) effects may also be important. These factors operate simultaneously and interact mutually so that precise assessment of their relative importance is difficult.  相似文献   

7.
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   

8.
The solution properties of fibrinogen and the thrombin-induced activation and gelation of fibrinogen in 95% D2O at pH 7.4 were compared to those in H2O under similar conditions. The initial release rates of fibrinopeptides A and B in D2O were slightly slower than those in H2O. However, the values of the Michaelis-Menten parameters Km and V for the release of the two peptides in D2O and H2O in the presence of 0.5 M NaCl were about the same. From turbidity measurements at 450 nm it is obvious that fibrinogen is soluble in a slightly more narrow range of NaCl concentration and that the fibrin gels have a higher degree of lateral aggregation in D2O than in H2O. The variation of fibrinogen concentration, thrombin concentration, pH and ionic a strength have a similar dependence on the final gel structure and clotting time in D2O and H2O. SDS-gel electrophoresis on fibrin samples, which were cross-linked by factor XIII, yielded results where the cross-linking of the gamma-chain appeared to be the same in D2O and H2O. The alpha-chain cross-linking was somewhat faster in D2O than in H2O. When fibrinogen solutions in 95% D2O were incubated at 20 mM CaCl2, a slow gelation of fibrinogen was observed, which was found to be induced by trace amounts of factor XIII. The final gel turbidity appeared to be about the same for this gelation as for that induced by thrombin. The differences in solubility for fibrinogen, kinetics for the enzyme reaction and optical properties for the fibrin gels in D2O and H2O may be explained by differences in electrostatic interactions, hydrogen bonding and hydration of fibrinogen in these two media.  相似文献   

9.
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   

10.
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   

11.
Isolated sheaths of Leptothrix discophora SP-6 (ATCC 51168) were tested for susceptibility to degradation by a variety of chemical denaturants and lytic enzymes and found to be resistant to many reagents and enzyme treatments. However, disulfide bond-reducing agents such as dithiothreitol (DTT), beta-mercaptoethanol, sodium cyanide, and sodium sulfite degraded the sheath, especially at elevated pH (pH 9) and temperature (50 degrees C). DTT and beta-mercaptoethanol caused more rapid degradation of the sheath than cyanide or sulfite. Treatment of the sheath with 1 N NaOH resulted in rapid breakdown, while treatment with 1 N HCl resulted in slow but significant hydrolysis. Transmission electron microscopy showed that the 6.5-nm fibrils previously shown to be an integral structural element of the sheath fabric (D. Emerson and W. C. Ghiorse, J. Bacteriol. 175:7808-7818, 1993) were progressively dissociated into random masses during DTT-induced degradation. Quantitation of disulfide bonds with DTT showed that the sheaths contained approximately 2.2 mumol of disulfides per mg of sheath protein. Reaction with 5,5'-dithio-bis-(2-nitrobenzoic acid) showed that sheaths also contained approximately 0.8 mumol of free sulfhydryls per mg of protein. A sulfhydryl-specific fluorescent probe (fluorescein 5-maleimide) showed that the free sulfhydryls in sheathed cell filaments were evenly distributed throughout the sheath. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis autoradiography of [14C]iodoacetamide-labeled sheaths and DTT-dissociated sheath fibril suspensions showed that the majority of 14C-labeled sulfhydryls in the sheaths did not enter the gel. However, low-molecular-mass silver-staining bands (14 to 45 kDa) did appear in the gels after iodoacetic acid or iodoacetamide alkylation of the dissociated fibrils. These bands did not stain with Coomassie blue. Their migration in gels was slightly affected by digestion with pronase. The fibrils contained 20 to 25% protein. These results confirm that the sheath fibrils consist of high molecular-weight heteropolysaccharide-protein complexes. We hypothesize that proteins in the fibril complexes provide interfibril cross-linking to maintain the structural integrity of the sheath.  相似文献   

12.
Fibrinogen-420 is a minor subclass of human fibrinogen that is so named because of its higher molecular weight compared to fibrinogen-340, the predominant form of circulating fibrinogen. Each of the two Aalpha chains of fibrinogen-340 is replaced in fibrinogen-420 by an Aalpha isoform termed alphaE. Such chains contain a globular C-terminal extension, alphaEC, that is homologous with the C-terminal regions of Bbeta and gamma chains in the fibrin D domain. The alphaEC domain lacks a functional fibrin polymerization pocket like those found in the D domain, but it does contain a binding site for beta2 integrins. Electron microscopy of fibrinogen-340 molecules showed the major core fibrinogen domains, D-E-D, plus globular portions of the C-terminal alphaC domains. Fibrinogen-420 molecules had two additional globular domains that were attributable to alphaEC. Turbidity measurements of thrombin-cleaved fibrinogen-420 revealed a reduced rate of fibrin polymerization and a lower maximum turbidity. Thromboelastographic measurements also showed a reduced rate of fibrin-420 polymerization (amplitude development) compared with fibrin-340. Nevertheless, the final amplitude (MA) and the calculated elastic modulus (G) for fibrin-420 were greater than those for fibrin-340. These results suggested a greater degree of fibrin-420 branching and thinner matrix fibers, and such structures were found in SEM images. In addition, fibrin-420 fibers were irregular and often showed nodular structures protruding from the fiber surface. These nodularities represented alphaEC domains, and possibly alphaC domains as well. TEM images of negatively shadowed fibrin-420 networks showed irregular fiber borders, but the fibers possessed the same 22.5-nm periodicity that characterizes all fibrin fibers. From this result, we conclude that fibrin-420 fiber assembly occurs through the same D-E interactions that drive the assembly of all fibrin fibrils, and therefore that the staggered overlapping molecular packing arrangement is the same in both types of fibrin. The alphaEC domains are arrayed on fiber surfaces, and in this location, they would very likely slow lateral fibril association, causing thinner, more branched fibers to form. However, their location on the fiber surface would facilitate cellular interactions through the integrin receptor binding site.  相似文献   

13.
beta-Amyloid protein (beta-A/4) is the major protein component of Alzheimer disease-related senile plaques and has been postulated to be a significant contributing factor in the onset and/or progression of the disease. In the senile plaque, beta-A/4 appears as bundles of amyloid fibrils. The biological activity of beta-A/4 may be related to its state of aggregation. In this work, self-assembly, fibril formation, and interfibrillary aggregation of beta(1-28), a synthetic peptide homologous with the amino-terminal fragment of beta-A/4, were investigated. The predominant form of beta(1-28) detected by size-exclusion chromatography and polyacrylamide gel electrophoresis was apparently a tetramer which does not bind Congo red. Aggregates containing cross-beta sheet structures which bind Congo red and thioflavin T were observed at concentrations of approximately 0.3 mg/ml or greater. Concentrations of 0.5-1 mg/ml were necessary for aggregation into fibrils to be detectable by classical or quasielastic light scattering. Both fibril elongation and fibril-fibril aggregation occur over the time scale investigated. The kinetics of aggregation were much faster at physiological salt concentrations than at lower ionic strength. Ionic strength also appeared to influence the morphology of the fibril aggregates. The data indicate that sample preparation method and sample history influence fibril size and number density.  相似文献   

14.
Kinetics of inhibition of fibrin monomer polymerization produced by Fab fragments prepared from immunochemically purified monospecific antibodies to the surface epitopes of different domains of fibrinogen molecule has been correlated with electron microscopic observations of resulting specimens. Fab fragments prepared from anti FgD antisera were the most efficient inhibitors of thrombin-catalysed conversion of fibrinogen to fibrin; polymerization of fibrin monomers as detected spectrophotometrically was abolished at 2:1 molar ratio of anti FgD Fab fragments to fibra monomer. These Fab fragments acting as a steric hindrance of polymerization sites inhibited the first stage of fibrin monomer aggregation. Interaction of Fab fragments derived from antibodies specific for alpha 239-476 with corresponding segment of fibrinogen molecule resulted in a weak inhibition of fibrin monomer polymerization. However, fibrin obtained in the presence of these Fab fragments was significantly modified and showed no periodicity. This observation may suggest that anti alpha 239-476 Fab impaired the course of the second stage of fibrin monomer polymerization, i.e. lateral association of fibrin fibrils.  相似文献   

15.
Fragment D (Mr 100 000) prepared from a terminal plasmin digest of fibrinogen was isolated and used to study its effect on fibrin formation. Increasing amounts of fragment D added to a solution of fibrinogen and thrombin decrease the rigidity of the resultant gel (10% of control at 2 mol of fragment D/mol of fibrinogen). Half-maximal inhibition is achieved at 1 mol of fragment D/mol of fibrinogen for non-cross-linked clots and at 1/2 mol of fragment D/mol of fibrinogen for cross-linked clots. "Clottability' decreases concomitantly with the rigidity. Only small amounts of fragment D (less than 10% for non-cross-linked gels) are incorporated into the gel. Light-scattering shows an increase in the final fibre thickness at fragment D concentrations up to 2 mol of fragment D/mol of fibrinogen, from 60 molecules/cross-section for the control to 120 molecules/cross-section. Higher fragment D concentrations lead to a decrease in the final fibre thickness. The limit fibre thickness is 8 nm, with a length of 80 nm, which is equivalent to a fibrin trimer. On the basis of results of synthetic-substrate and fibrinopeptide-release assays, it is clear that thrombin inactivation is not responsible for this effect. These data suggest that fragment D may inhibit fibrin formation by blocking the bimolecular polymerization of activated fibrin monomer molecules to form protofibrils, although additional effects on subsequent assembly steps may also be involved.  相似文献   

16.
The effect of fibrinogen on the two steps of polymerization of two fibrin forms differing in the set of polymerization sites (fibrin-desAA and fibrin-desAABB) was studied. It was shown that fibrinogen inhibited the protofibril growth and fibril formation at the stage of lateral aggregation more effectively with fibrin-desAABB than with fibrin desAA. When the fibrinogen D2-site was blocked by tetrapeptide Gly-His-Arg-Pro, the key structure of the E2-site, the inhibitory activity of fibrinogen diminished. A conclusion is drawn that the high susceptibility of fibrin-desAABB to fibrinogen is due to the interaction of the E2-active site with the D2-site of the fibrinogen molecule. The concentration dependence of the tetrapeptide Gly-His-Arg-Pro-induced inactivation of fibrinogen and the effects of temperature and Ca2+ on the tetrapeptide interaction with fibrinogen were investigated.  相似文献   

17.
The carbohydrate moiety of some glycoproteins influences their secretion and functional properties. We have examined the importance of the oligosaccharide chains of fibrinogen in this regard. Fibrinogen was labeled de novo by the addition to rabbit hepatocyte monolayer cultures of either 3H-amino-acids or [2-3H] mannose, in the presence or absence of tunicamycin, a potent inhibitor of glycosylation. Inhibition of glycosylation, which ranged from 75 to 80%, was determined by incorporation of [2-3H]mannose as quantitated by gel filtration. Synthesis and secretion of fibrinogen were quantitated by 3H-amino-acid incorporation, using anti-fibrinogen immunoaffinity column chromatography of medium and cell homogenates. Tunicamycin did not appreciably inhibit fibrinogen synthesis, as compared to a 30-40% inhibition of overall protein synthesis, determined by incorporation of 3H-amino-acids into trichloroacetic acid-precipitable material. There was no evidence that secretion of fibrinogen was impaired. Fibrinogen from medium was copurified by adding cold plasma fibrinogen as carrier. Nonglycosylated fibrinogen was found to be functional as demonstrated by incorporation of radioactivity into clots of the copurified material at a rate identical to that of glycosylated fibrinogen. When clotted in the presence of Ca2+ and Factor XIII, cross-linking of glycosylated and nonglycosylated fibrin was demonstrable on fluorography of sodium dodecyl sulfate-polyacrylamide gels, showing disappearance of gamma-chain and appearance of gamma-gamma-dimers.  相似文献   

18.
The structure of proteolytically modified fibrin and a closely related modified fibrinogen aggregate have been studied by analysis of electron microscope images. For both structures, we propose a model that consists of double-stranded, 2-fold helical protofibrils, which are associated laterally to form ordered fibrils, with a C222 space group: a = 44.0 nm, b = c = 9.4 nm. Each fibril is 80 nm or less in diameter, and twists along its length in a right-handed sense, with a pitch from 7 to 12 times the molecular length. The fibrils associate laterally to form bundles, which tend to twist in a left-handed sense, with a pitch of the order of 40 times the molecular length. The specific volume of modified fibrin calculated from this model is 3.9 A3 per dalton, which is comparable to the specific volume of 3.6 A3 per dalton for modified fibrinogen crystals but is lower than the 6 A3 per dalton determined for fibrin from light-scattering experiments. Comparison of our electron microscope results with X-ray and neutron diffraction data suggest a similar, but less well-ordered, structure for native fibrin, with a smaller fibril, approximately 18.4 nm wide, consisting of eight protofibrils.  相似文献   

19.
Surface properties of fibrinogen and fibrin   总被引:1,自引:0,他引:1  
By contact angle measurements on layers of fibrinogen and fibrin, it can be shown that the transformation from fibrinogen to fibrin is accompanied by a change in surface properties from very hydrophilic (fibrinogen) to moderately but definitely hydrophobic (fibrin). It is also shown that, contrary to serum albumin and gamma globulin, fibrinogen does not become more hydrophobic upon drying.  相似文献   

20.
Fibrin is a biopolymer that gives thrombi the mechanical strength to withstand the forces imparted on them by blood flow. Importantly, fibrin is highly extensible, but strain hardens at low deformation rates. The density of fibrin in clots, especially arterial clots, is higher than that in gels made at plasma concentrations of fibrinogen (3–10 mg/mL), where most rheology studies have been conducted. Our objective in this study was to measure and characterize the elastic regimes of low (3–10 mg/mL) and high (30–100 mg/mL) density fibrin gels using shear and extensional rheology. Confocal microscopy of the gels shows that fiber density increases with fibrinogen concentration. At low strains, fibrin gels act as thermal networks independent of fibrinogen concentration. Within the low-strain regime, one can predict the mesh size of fibrin gels by the elastic modulus using semiflexible polymer theory. Significantly, this provides a link between gel mechanics and interstitial fluid flow. At moderate strains, we find that low-density fibrin gels act as nonaffine mechanical networks and transition to affine mechanical networks with increasing strains within the moderate regime, whereas high-density fibrin gels only act as affine mechanical networks. At high strains, the backbone of individual fibrin fibers stretches for all fibrin gels. Platelets can retract low-density gels by >80% of their initial volumes, but retraction is attenuated in high-density fibrin gels and with decreasing platelet density. Taken together, these results show that the nature of fibrin deformation is a strong function of fibrin fiber density, which has ramifications for the growth, embolization, and lysis of thrombi.  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