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 0 毫秒
1.
R Procyk  R G King 《Biopolymers》1990,29(3):559-565
The elastic modulus (G') of factor XIIIa induced fibrinogen gels was found to be substantially lower than the G' of fibrin gels that were formed by clotting fibrinogen with thrombin. The addition of fibronectin and/or the reducing reagent dithiothreitol (DTT) to the factor XIIIa coagulation mixture led to the formation of a weaker gel structure, while the rigidity of thrombin induced clots was not appreciably affected by the inclusion of the DTT but increased somewhat in the presence of fibronectin. The reasons for the differing clot rigidities are discussed in terms of biochemical mechanisms.  相似文献   

2.
The human fibrinogen gamma-chain C-terminal segment functions as the platelet integrin binding site as well as the Factor XIIIa cross-linking substrate and thus plays an important role in blood clot formation and stabilization. The three-dimensional structure of this segment has been determined using carrier protein driven crystallization. The C-terminal segment, gamma-(398-411), was attached to a linker sequence at the C-terminus of glutathione S-transferase and the structure of this fusion protein determined at 1.8 A resolution. Functional studies of the chimeric protein demonstrate that the fibrinogen sequence in the presence of the carrier protein retains its specific functions as ligand for platelet integrin alpha(IIb)beta3 (gpIIb/IIIa) and as a cross-linking substrate for Factor XIIIa. The structure obtained for the fibrinogen gamma-chain segment is not affected by crystal packing and can provide the missing links to the recently reported model of cross-linked fibrin.  相似文献   

3.
The complex modulus (E*) and elastic modulus (E') of agarose gels (2% to 4%) are measured with a dynamic mechanical analyzer in frequency sweep shear sandwich mode between 0.1 and 20 Hz. The data showed that E* and E' increase with frequency according to a power law which can be described by a fractional derivative model to characterize the dynamic viscoelasticity of the gel. The functions between the model parameters including storage modulus coefficient (H) and the power law exponent (beta) and the agarose concentration are established. A molecular basis for the application of the fractional derivative model to gel polymers is also discussed. Such an approach can be useful in tissue culture studies employing dynamic pressurization or for validation of magnetic resonance elastography.  相似文献   

4.
Regulation of formation of factor XIIIa by its fibrin substrates   总被引:6,自引:0,他引:6  
S D Lewis  T J Janus  L Lorand  J A Shafer 《Biochemistry》1985,24(24):6772-6777
Thrombin-catalyzed release of activation peptide (AP) from plasma factor XIII was studied to characterize the regulation of this initial step in the activation of factor XIII zymogen (fibrin-stabilizing factor). High-performance liquid chromatography was used to monitor the kinetics of release of AP. Non-cross-linked polymeric fibrins I and II (polymerized des-A- and des-A,B-fibrinogens), physiological substrates of factor XIIIa, were shown to be potent promoters of thrombin-catalyzed release of activation peptide from factor XIII. These promoters are proposed to act by complexing factor XIII and reducing the apparent Km for thrombin-catalyzed release of AP. Since thrombin-catalyzed release of AP is inefficient in the absence of polymerized fibrin, this mode of regulation should minimize formation of factor XIIIa prior to the formation of its fibrin substrates. The promoting activity of polymeric fibrin was rapidly lost when catalytically competent factor XIIIa was allowed to form. This observation suggested the possibility that factor XIIIa catalyzed cross-linking of fibrin inactivates fibrin as a promoter for the thrombin-catalyzed release of AP from factor XIII. Consistent with this view, the thiol reagent S-methyl methanethiosulfonate inactivated factor XIIIa, blocked cross-linking of fibrin, and protected against loss of its promoter activity. This mode of feedback regulation of the activation process by catalytically active factor XIIIa may serve to ensure against continued generation of factor XIIIa after its fibrin substrates have been cross-linked.  相似文献   

5.
We have designed and developed a microfluidic system to study the response of cells to controlled gradients of mechanical stiffness in 3D collagen gels. An 'H'-shaped, source-sink network was filled with a type I collagen solution, which self-assembled into a fibrillar gel. A 1D gradient of genipin--a natural crosslinker that also causes collagen to fluoresce upon crosslinking--was generated in the cross-channel through the 3D collagen gel to create a gradient of crosslinks and stiffness. The gradient of stiffness was observed via fluorescence. A separate, underlying channel in the microfluidic construct allowed the introduction of cells into the gradient. Neurites from chick dorsal root ganglia explants grew significantly longer down the gradient of stiffness than up the gradient and than in control gels not treated with genipin. No changes in cell adhesion, collagen fiber size, or density were observed following crosslinking with genipin, indicating that the primary effect of genipin was on the mechanical properties of the gel. These results demonstrate that (1) the microfluidic system can be used to study durotactic behavior of cells and (2) neurite growth can be directed and enhanced by a gradient of mechanical properties, with the goal of incorporating mechanical gradients into nerve and spinal cord regenerative therapies.  相似文献   

6.
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   

7.
A longstanding challenge in accurate mechanical characterization of engineered and biological tissues is maintenance of both stable sample hydration and high instrument signal resolution. Here, we describe the modification of an instrumented indenter to accommodate nanomechanical characterization of biological and synthetic tissues in liquid media, and demonstrate accurate acquisition of force-displacement data that can be used to extract viscoelastoplastic properties of hydrated gels and tissues. We demonstrate the validity of this approach via elastoplastic analysis of relatively stiff, water-insensitive materials of elastic moduli E>1000 kPa (borosilicate glass and polypropylene), and then consider the viscoelastic response and representative mechanical properties of compliant, synthetic polymer hydrogels (polyacrylamide-based hydrogels of varying mol%-bis crosslinker) and biological tissues (porcine skin and liver) of E<500 kPa. Indentation responses obtained via loading/unloading hystereses and contact creep loading were highly repeatable, and the inferred E were in good agreement with available macroscopic data for all samples. As expected, increased chemical crosslinking of polyacrylamide increased stiffness (E40 kPa) and decreased creep compliance. E of porcine liver (760 kPa) and skin (222 kPa) were also within the range of macroscopic measurements reported for a limited subset of species and disease states. These data show that instrumented indentation of fully immersed samples can be reliably applied for materials spanning several orders of magnitude in stiffness (E=kPa-GPa). These capabilities are particularly important to materials design and characterization of macromolecules, cells, explanted tissues, and synthetic extracellular matrices as a function of spatial position, degree of hydration, or hydrolytic/enzymatic/corrosion reaction times.  相似文献   

8.
The kinetics of activation of platelet factor XIII, an a-subunit dimer, were characterized by determining rate constants for activation peptide (AP) release, generation of activity, and exposure of the active-site thiol group. The specificity constant (kappacat/Km) for alpha-thrombin-catalyzed AP release, 1.2 x 10(5) M-1s-1, was found to be similar to that for AP release from the tetramer plasma factor XIII (a2b2) [Janus, T.J., Lewis, S. D., Lorand, L., & Shafer, J. A. (1983) Biochemistry 22, 6269-6272], implying that the b subunits of plasma factor XIII do not hinder alpha-thrombin-catalyzed cleavage of AP from the a subunit. Platelet factor XIIIa activity was generated at a rate approximately twice the rate of AP release. This difference in rates was shown to be consistent with a reaction pathway for activation of platelet factor XIII wherein full factor XIIIa activity is generated when one AP is removed from the dimeric zymogen so that removal of the second AP has no detectable effect on catalytic activity. In accord with this conclusion, the rate constant for exposure of the active-site thiol group, as measured by the incorporation of [1-14C]-iodoacetamide, was about twice that observed for the removal of AP.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
C Tharaud  A M Ribet  C Costes  C Gaillardin 《Gene》1992,121(1):111-119
The industrial yeast, Yarrowia lipolytica, secretes high yields of an alkaline extracellular protease (AEP), which is synthesized as a preproprotein encoded by the XPR2 gene. We investigated the possibility of using this system for the secretion of human coagulation factor XIII subunit a (FXIIIa). This protein is naturally secreted in the plasma by an unknown, signal peptide-independent mechanism and has so far been found to be nonsecretable in yeast. We have designed six hybrid genes encoding fusion proteins between increasing portions of the AEP preprodomain and the precursor or mature forms of FXIIIa. All constructs directed translocation of the FXIIIa precursor into the endoplasmic reticulum. Transport of the translocated and core-glycosylated hybrid precursor to the Golgi apparatus appeared to be strongly rate limiting, and most of the precursors appeared to be partially proteolysed. One of these constructs directed the extracellular secretion of a low amount of hyperglycosylated FXIIIa. These results indicate that fusion to the yeast AEP signal peptide and dipeptide stretch allows FXIIIa to be translocated, albeit inefficiently, through the endoplasmic reticulum and to follow a classical secretory transit.  相似文献   

10.
The dansyl hapten in dansylcadaverine offers unique possibilities for measuring the incorporation of the monoamine into proteins (e.g. N,N'-dimethylcasein) by transamidating enzymes such as factor XIIIa and the transglutaminases. The protein-bound dansylcadaverine was assayed by an ELISA procedure based on a monoclonal antibody to the dansyl moiety.  相似文献   

11.
Human plasma fibrin stabilizing factor (factor XIII) may be separated from fibrinogen through reversible fibrinogen polymer formation at pH 6.6, gamma/2 0.3, 0 degrees C, and subsequent Bio-Gel A 1.5m filtration. Factor XIII activity is eluted after the monomer fibrinogen peak. Polymer fractions from eight preparations, processed in duplicate, contain a mean 0.002 units factor XIII per mg fibrinogen, or about 0.7% the factor XIII content of standard plasma. Factor XIII-free fibrinogen polymers are easily dissociated (greater than 98%) to the monomer form by incubation at 37 degrees C, 18 hours. The fibrinogen preparations utilized were devoid of plasma fibronectin; thus these studies also show that reversible human fibrinogen polymer formation occurs in its absence.  相似文献   

12.
Using a multitechnique approach, two temperature domains have been identified in agarose gelation. Below 35 degrees C, fast gelation results in strong, homogeneous and weakly turbid networks. The correlation length, evaluated from the wavelength dependence of the turbidity, is close to values of pore size reported in the literature. Above 35 degrees C, gelation is much slower and is associated with the formation of large-scale heterogeneities that can be monitored by a marked change in the wavelength dependence of turbidity and visualised by transmission electron microscopy. Curing agarose gels at temperatures above 35 degrees C, and then cooling them to 20 degrees C, produces much weaker gels than those formed directly at 20 degrees C. Dramatic reductions in the elastic modulus and failure strain and stress are found in this case as a result of demixing during cure. An interpretation, based on the kinetic competition between osmotic forces (in favor of phase separation) and elastic forces (that prevent it) is proposed.  相似文献   

13.
We studied the binding of 125I-platelet and plasma Factor XIII (125I-Factor XIII) to human platelets. When 125I-Factor XIII was incubated with gel-filtered platelets, calcium chloride (5 mM) and thrombin (1 unit/ml) at 37 degrees C, saturable binding was observed. Half-maximal binding occurred at 1 min. Binding was inhibited 93% by a 100-fold molar excess of unlabeled ligand but not by other purified proteins. Greater than 87% of platelet-bound radioactivity migrated as thrombin-cleaved a-chains (a'-chains) in sodium dodecyl sulfate-polyacrylamide gels indicating that Factor XIIIa but not Factor XIII binds to platelets. 125I-Factor XIIIa does not bind to unstimulated platelets. When platelet secretion was blocked, binding was markedly inhibited. 125I-Factor XIIIa bound minimally to platelets stimulated with agonists other than thrombin. Thus, binding is dependent on platelet activation, as well as modification of platelets by thrombin. 125I-Factor XIIIa bound to gamma-thrombin-stimulated platelets, at concentrations which did not clot fibrinogen. Therefore, Factor XIIIa is not bound to fibrin associated with platelets. Binding was only partially reversible. Approximately 12,000 molecules of Factor XIIIa were bound per platelet. 125I-Factor XIIIa bound normally to platelets from patients with severe Glanzmann's thrombasthenia indicating that 125I-Factor XIIIa does not bind to platelet glycoproteins IIb or IIIa, or platelet-bound fibrinogen. Chymotrypsin treatment of platelets inhibited 125I-Factor XIIIa binding by 78% without inhibiting secretion. Methylamine and putrescine, Factor XIIIa substrates, and N-ethylmaleimide, an active site inhibitor, did not inhibit binding. Factor XIIIa bound to platelets was enzymatically active and catalyzed [3H]putrescine incorporation into platelet proteins. The specific binding of Factor XIIIa to platelets suggests it may play a role in physiologic reactions involving platelets.  相似文献   

14.
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   

15.
The in vitro and in vivo functionality of the anionic plant polysaccharide pectin depends not only on the amount of ion-binding groups attached to the polymer but also on the distribution of such groups along the backbone. It has been proposed recently that information regarding this intramolecular distribution can be quantified by defining a degree of blockiness (DB or DB(abs)), and the usefulness of such measures in discriminating qualitatively between pectins originating from different sources has been demonstrated. Despite this, the value of these parameters in predicting the pseudoequilibrium elastic modulus of gels remains untested. This study seeks to address this problem through the sourcing and in-house modification of a variety of pectins in order to produce a library of distinct representative fine structures. These were subsequently characterized in terms of their relevant properties, including the determination of the proposed DB and DB(abs), and the formation of gels of these samples was monitored using small deformation mechanical spectroscopy. In addition to ionotropic calcium gels the effect of the fine structure on acid gelation was also studied.  相似文献   

16.
The diffusive properties of biomacromolecules within the aqueous phase of polyacrylamide gels are described. High quality NMR spectra can be obtained under such conditions. As compared to water, a fivefold reduction in the translational diffusion constant, but only a 1.6-fold decrease (1.4-fold increase) in amide-15N T2 (T1) are observed for human ubiquitin within a 10% acrylamide gel. Weak alignment of the solute macromolecules can be achieved within such gels by vertical or radial compression or by the embedding of magnetically oriented purple membrane fragments. The methods are applied to derive residual dipolar couplings for human HIV-1 Nef and ubiquitin.  相似文献   

17.
The proteases of several species of African trypanosomes were analysed by electrophoresis in sodium dodecyl sulphate/polyacrylamide gels containing fibrinogen or collagen. After electrophoresis the gels were incubated in the presence of enzyme activators and/or inhibitors and then stained with Coomassie brilliant blue. The areas where the proteolytic activity had degraded the fibrinogen did not stain and so formed clear bands against a blue background. The proteases were found to have pH optima between 5 and 6, and required dithiothreitol or 2-mercaptoethanol for full expression of their activity. They were inhibited by amino acid chloromethanes, iodoacetamide, p-chloromercuribenzoate and other inhibitors of the thiol-dependent proteases, as well as by the trypanocidal drugs berenil (4,4'-diamidinodiazoaminobenzene-diacetamidoacetate) and pentamidine [1,5-di-(4-amidinophenoxy)pentane-di-(2- hydroxyethanesulphonate)]. Trypanosoma evansi, Trypanosoma brucei brucei and Trypanosoma brucei gambiense each have a protease with a relative molecular mass, Mr, of 28 000. In addition they occasionally exhibit activity at higher Mr values (up to 105000). Trypanosoma congolense has a low-Mr protease (31 000) and may exhibit higher-Mr proteases (up to 150000). The protease profiles of Trypanosoma vivax differ from the other species, T. brucei or T. congolense, and are present in lesser amounts. The proteases of the cultured procyclic forms are present in much smaller amounts than those of the metacyclic or mammalian blood stream forms of these parasites. The catalytic properties and inhibition characteristics of these thiol-dependent enzymes suggest that they resemble the mammalian lysosomal cathepsins B and L.  相似文献   

18.
19.
Blood coagulation factor XIIIa is a calcium-dependent enzyme that covalently ligates fibrin molecules during blood coagulation. X-ray crystallography studies identified a major calcium-binding site involving Asp(438), Ala(457), Glu(485), and Glu(490). We mutated two glutamic acid residues (Glu(485) and Glu(490)) and three aspartic acid residues (Asp(472), Asp(476), and Asp(479)) that are in close proximity. Alanine substitution mutants of these residues were constructed, expressed, and purified from Escherichia coli. The K(act) values for calcium ions increased by 3-, 8-, and 21-fold for E485A, E490A, and E485A,E490A, respectively. In addition, susceptibility to proteolysis was increased by 4-, 9-, and 10-fold for E485A, E490A, and E485A,E490A, respectively. Aspartic acids 472, 476, and 479 are not involved directly in calcium binding since the K(act) values were not changed by mutagenesis. However, Asp(476) and Asp(479) are involved in regulating the conformation for exposure of the secondary thrombin cleavage site. This study provides biochemical evidence that Glu(485) and Glu(490) are Ca(2+)-binding ligands that regulate catalysis. The binding of calcium ion to this site protects the molecule from proteolysis. Furthermore, Asp(476) and Asp(479) play a role in modulating calcium-dependent conformational changes that cause factor XIIIa to switch from a protease-sensitive to a protease-resistant molecule.  相似文献   

20.
1. The activation of the native enzyme was achieved by a proteolytic procedure involving thrombin. 2. The pH profile was independent of the nature of the substrates assayed (casein or dimethylcasein plus putrescine). The optimum pH was between 7.6 and 7.9 and the pK values were 6.5/7 and 8.7/9. A cysteinyl residue appeared to be involved in the pH-dependence activity. 3. In the presence of calcium, the thermostability of enzyme was increased: the temperature at which enzyme lost half of its activity increased up to 7 degrees C. 4. The kinetics of the thermal deactivation of F XIIIa depended on the presence or absence of calcium. 5. In its presence the reaction obeyed second order kinetics, while in its absence, the kinetics were of first order. In the first case, the irreversible thermal deactivation could be described by a two-step mechanism (N----X----D) while in the second case, the deactivation followed the simple model (N----D). 6. Neither divalent cations like Sr2+, Ba2+, Mg2+, nor bovine serum-albumin and polyhydric alcohols were able to increase the thermostability of F XIIIa. 7. Thermal deactivation of F XIIIa did not appear linked to the redox state of enzyme, nor to the modification of SH groups. 8. We observed a good correlation between the loss of activity and the unfolding of the polypeptide chain of F XIIIa during heating. 9. The optimum temperature of F XIIIa activity was 40 degrees C at pH 8 and 45 degrees C at pH 7.  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