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 15 毫秒
1.
The interactions of two proteinase inhibitors, heparin cofactor II and antithrombin, with thrombin are potentiated by heparin. Using two methods, we have studied the potentiating effects of a series of heparin (poly)saccharides with high affinity for antithrombin and mean Mr ranging from approx. 1700 to 18,800. First, catalytic amounts of heparin (poly)saccharide were added to purified systems containing thrombin and either heparin cofactor II or antithrombin. Residual thrombin activity was determined with a chromogenic substrate. It was found that only the higher-Mr polysaccharides (Mr greater than 8000) efficiently catalysed thrombin inhibition by heparin cofactor II, there being a progressive catalytic effect with increasing Mr of the polysaccharide. Weak accelerating effects were noted with low-Mr saccharides (Mr less than 8000). This contrasted with the well-characterized interaction of heparin with antithrombin and thrombin, where heparin oligosaccharides of Mr less than 5400 had absolutely no ability to accelerate the reaction, while (poly)saccharides of Mr exceeding 5400 showed rapidly increasing catalytic activity with increasing Mr. Secondly, these and other heparin preparations were added in a wide concentration range to plasma with which 125I-labelled thrombin was then incubated for 30 s. Inhibited thrombin was determined from the distribution of labelled thrombin amongst inhibitor-thrombin complexes, predominantly antithrombin-thrombin and heparin cofactor II-thrombin complexes. In this situation, where the inhibitors competed for thrombin and for the (poly)saccharides, it was found that, provided the latter were of high affinity for antithrombin and exceeded a Mr of 5400, thrombin inhibition in plasma was mediated largely through antithrombin. Polysaccharides of Mr exceeding 8000 that were of low affinity for antithrombin accelerated thrombin inhibition in plasma through their interaction with heparin cofactor II. High concentrations of saccharides of Mr 1700-5400 exhibited a size-dependent acceleration of thrombin inhibition, not through their interaction with antithrombin, but through their interaction with heparin cofactor II.  相似文献   

2.
The anisotropy of the fluorescence of dansyl (5-dimethylaminonaphthalene-1- sulphonyl) groups covalently attached to human platelet factor 4 was used to detect the macromolecular compounds formed when the factor was mixed with heparin. At low heparin/protein ratios a very-high-molecular-weight compound (1) was formed that dissociated to give a smaller compound (2) when excess heparin was added. 2. A large complex was also detected as a precipitate that formed at high protein concentrations in chloride buffer. It contained 15.7% (w/w) polysaccharide, equivalent to four or five heparin tetrasaccharide units per protein tetramer. In this complex, more than one molecule of protein binds to each heparin molecule of molecular weight greater than about 6 X 10(3).3. The stability of these complexes varied with pH, salt concentration and the chain length of the heparin. The limit complexes found in excess of the larger heparins consisted of only one heparin molecule per protein tetramer, and the failure to observe complexes with four heparin molecules/protein tetramer is discussed.  相似文献   

3.
Heparin with high anticoagulant activity was isolated from the two marine clam species Anomalocardia brasiliana and Tivela mactroides. A large portion of the polysaccharide chains of both preparations bound with high affinity to immobilized antithrombin. Titrations monitored by tryptophan fluorescence showed that clam polysaccharide chains with Mr approximately 22,500 contained up to three binding sites for antithrombin and that the binding constants for the interaction of these chains with antithrombin were higher than those reported for mammalian heparin of comparable size. Structural analysis of clam heparin fractions and subfractions of clam heparin with differing affinity for immobilized antithrombin revealed the presence of large amounts (up to 25-30% of the total disaccharide units) of the 3-O-sulfated saccharide sequences (-GlcNSO3)-GlcA-GlcNSO3(3-OSO3)- and (-GlcNSO3)-GlcA-GlcNSO3(3,6-di-OSO3)-, previously identified as unique markers for the antithrombin-binding region of heparin. The content of these saccharide sequences was found to increase with increasing affinity of the parent polysaccharide for antithrombin. Structural analysis of the clam heparins also demonstrated the occurrence of a novel saccharide sequence, tentatively identified as (-GlcNSO3)-IdA-GlcNSO3(3,6-di-OSO3)-, that has not previously been found in heparin or related polysaccharides. The contents of this latter sequence, at most 3-4% of the total disaccharide units, showed no correlation with the affinity for antithrombin.  相似文献   

4.
Kinetic characteristics of several heparin preparations and substitute heparins were determined to help understand the bases for activity differences. Several materials were highly active in factor Xa inhibition and the reaction rate at constant factor Xa concentration appeared to be predicted by the extent of intrinsic antithrombin III fluorescence change induced by the polysaccharide. Heparin fractions of different molecular weight and affinity for antithrombin III showed similar kinetic parameters in catalysis of the thrombin-antithrombin III reaction when these parameters were expressed on the basis of antithrombin III-binding heparin. The latter was determined by stoichiometric titration of the antithrombin III fluorescence change by the heparin preparation. However, the various heparin fractions showed very different specific activities per mg of total polysaccharide. This indicated that functional heparin molecules had similar kinetic properties regardless of size or antithrombin III-binding affinity and is possible because the Km for antithrombin III is determined by diffusion rather than by binding affinity. Substitute heparins and depolymerized heparin were poor catalysts for thrombin inhibition, due at least partially to their affinity for thrombin. This latter binary interaction inhibits thrombin reaction in the heparin-catalyzed reaction.  相似文献   

5.
Small-angle neutron scattering was used to confirm that human platelet factor 4 was a compact tetrameric globular protein of radius of gyration 1.74 nm and indistinguishable from a sphere. The same technique, when applied to the 1:1 mol/mol complex of platelet factor and heparin of Mr 14000, revealed that the radius of gyration of the particle varied, depending on the relative proportion of 2H2O to H2O in the solvent. Analysis of this variation by the method of Ibel and Stuhrmann (Ibel, K. and Stuhrmann, H.B. (1975) J. Mol. Biol. 93, 255-266) revealed that in the complex the material of greatest neutron-scattering length (the highly sulphated polysaccharide heparin) was furthest from the centre of the particle. This confirms the postulate of Luscombe and Holbrook (Luscombe, M. and Holbrook, J.J. (1983) in Glycoconjugates (Chester, A.M., Heineg?rd, D., Lundblad, A. and Svensson, S., eds.), pp. 818-819, Secretariat, Lund) that the exact 1:1 mole ratio of heparin (Mr greater than 10 000) to platelet factor in this stable complex arises from the heparin winding around the outside of a globular protein core.  相似文献   

6.
The interference of the heparin-neutralizing plasma component S protein (vitronectin) (Mr = 78,000) with heparin-catalyzed inhibition of coagulation factor Xa by antithrombin III was investigated in plasma and in a purified system. In plasma, S protein effectively counteracted the anticoagulant activity of heparin, since factor Xa inhibition was markedly reduced in comparison to heparinized plasma deficient in S protein. Using purified components in the presence of heparin, S protein induced a concentration-dependent reduction of the inhibition rate of factor Xa by antithrombin III. This resulted in a decrease of the apparent pseudo-first order rate constant by more than 10-fold at a physiological ratio of antithrombin III to S protein. S protein not only counteracted the anticoagulant activity of commercial heparin but also of low molecular weight forms of heparin (mean Mr of 4,500). The heparin-neutralizing activity of S protein was found to be mainly expressed in the range 0.2-10 micrograms/ml of high Mr as well as low Mr heparin. S protein and high affinity heparin reacted with apparent 1:1 stoichiometry to form a complex with a dissociation constant KD = 1 X 10(-8) M as determined by a functional assay. As deduced from dot-blot analysis, direct interaction of radiolabeled heparin with S protein revealed a dissociation constant KD = 4 X 10(-8) M. Heparin binding as well as heparin neutralization by S protein increased significantly when reduced/carboxymethylated or guanidine-treated S protein was employed indicating the existence of a partly buried heparin-binding domain in native S protein. Radiolabeled heparin bound to the native protein molecule as well as to a BrCN fragment (Mr = 12,000) containing the heparin-binding domain as demonstrated by direct binding on nitrocellulose replicas of sodium dodecyl sulfate-polyacrylamide gels. Kinetic analysis revealed that the heparin neutralization activity of S protein in the inhibition of factor Xa by antithrombin III could be mimicked by a synthetic tridecapeptide from the amino-terminal portion of the heparin-binding domain. These data provide evidence that the heparin-binding domain of S protein appears to be unique in binding to heparin and thereby neutralizing its anticoagulant activity in the inhibition of coagulation factors by antithrombin III. The induction of heparin binding and neutralization may be considered a possible physiological mechanism initiated by conformational alteration of the S protein molecule.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

7.
The tributylammonium salt of a porcine heparin subfraction with low affinity for antithrombin III (Mr 7,500-18,000; anti-clotting activity, 7 USP units/mg), having degrees of sulfate substitution at D-glucosamine and L-iduronic acid residues of GlcNS 0.786, GlcN6s 0.628, and IdoA2s 0.682 mol, was reacted with 10 or 20 mol of pyridine-sulfur trioxide per mol equiv. of available hydroxyl groups in N,N-dimethylformamide at -10 degrees C for 1 h. Both chemical and NMR spectroscopic analyses revealed that sulfation proceeded exclusively at HO-6 in D-glucosamine and HO-2 in L-iduronic acid residues, according to the amount of the sulfating reagent used (GlcNS: 0.825, 0.830; GlcN6s: 0.872, 0.928; IdoA2s: 0.687, 0.749 mol, respectively). Affinity chromatography of the sulfated products on antithrombin III-Sepharose gel indicated that the polysaccharide acquired some affinity for the protein following the sulfation, as shown by the increase in the proportion of the high-affinity heparin fraction (%) from 1.1 to 6.7. Biological examination of these products indicated that sulfation at natural positions along with the polysaccharide chain resulted in significant increases in all the activities (blood anti-clotting, anti-Factor IIa, and anti-Factor Xa), and in the strength of intrinsic fluorescence of antithrombin III.  相似文献   

8.
Rat skin heparin proteoglycan labelled biosynthetically with 35S was fractionated on a column of antithrombin-Sepharose into fractions with varying degrees of affinity for antithrombin. These were treated with NaOH to release heparin chains (Mr 60,000-100,000), by beta-elimination or incubated with serum to produce fragments of the same order of size as commercial heparin (Mr 5000-30,000), by endoglycosidase cleavage. Chains and fragments were then fractionated on antithrombin-Sepharose. The various fractions were deaminated with HNO2 at pH 1.5 followed by reduction with NaB3H4. Approx 90% of the incorporated 3H was associated with disaccharides. These were fractionated by high-performance ion-exchange chromatography. A unique minor component corresponding to the sequence glucuronosyl-N-sulphoglucosaminyl (3,6-di-O-sulphate) in the polysaccharide was found only in fractions with high affinity for antithrombin. The glucosamine residue linked to C-4 of this glucuronosyl unit was predominantly (or exclusively) N-sulphated rather than N-acetylated, pointing to a structural difference between the antithrombin-binding region of rat heparin and that of pig mucosal heparin. Calculations based on the distribution of the glucosaminyl 3-O-sulphate group showed that approximately two-thirds of the total antithrombin-binding regions present in the unfractionated material were accommodated by only 20% of the proteoglycan molecules, and by 10% of the polysaccharide chains. While most of the proteoglycan molecules thus lacked such regions (and hence affinity for antithrombin) a minor proportion of the polysaccharide chains contained on the average three binding regions per molecule. These findings support by direct chemical analysis an earlier proposal, based on anticoagulant activities of similar rat skin heparin fractions, that the distribution of antithrombin-binding sites in intact heparin proteoglycans is markedly non-random.  相似文献   

9.
Fluorescent perylenediimide probes modified with 2, 4, 6, or 8 ammonium groups were synthesized and their binding to the antithrombotic drug heparin was studied by fluorescence spectroscopy in solution. The polyanionic polysaccharide strands of heparin bind more probe molecules per sugar unit when the charge of the latter is low, and stability of the probe-heparin complex increases with increasing probe charge.  相似文献   

10.
Pulse-labelling of mouse mastocytoma cell cultures, established from ascites fluid, with inorganic [35S]sulphate for 1 h yielded labelled heparin proteoglycan containing polysaccharide chains of Mr 60,000-100,000. After chase incubation for 24 h most of the 35S appeared in intracellular polysaccharide fragments similar in size to commercially available heparin, Mr 5000-25,000, as indicated by gel chromatography. Products isolated from cultures after 6 h of chase incubation consisted of partially degraded free polysaccharide chains and, in addition, residual proteoglycans that were of smaller size than the proteoglycans initially pulse-labelled. The polysaccharide chains released by alkali treatment from the residual chase-incubated proteoglycans were of the same size as the chains derived from proteoglycans after 1 h of pulse labelling. These results suggest that the intracellular degradation of heparin proteoglycan to polysaccharide fragments is initiated by release of intact polysaccharide chains, probably by action of a peptidase, and is pursued through cleavage of these chains by an endoglycosidase. An endoglucuronidase with stringent substrate specificity [Thunberg, Bäckström, Wasteson, Ogren & Lindahl (1982) J. Biol. Chem. 257, 10278-10282] has previously been implicated in the latter step. Cultures of more purified mastocytoma cells (essentially devoid of macrophages) did not metabolize [35S]heparin proteoglycan to polysaccharide fragments, but instead accumulated free intact polysaccharide chains, i.e. the postulated intermediate of the complete degradation pathway. When such purified cells were co-cultured with adherent mouse peritoneal cells, presumably macrophages, formation of polysaccharide fragments was observed. It is tentatively proposed that the expression of endoglucuronidase activity by the mast cells depends on collaboration between these cells and macrophages.  相似文献   

11.
K5 lyase of coliphage K5A degrades the K5 polysaccharide of encapsulated E. coli strains expressing the K5 antigen thereby contributing to virus binding and infection. We have investigated the affinities of the recombinant enzyme for different GAG ligands by isothermal fluorescence titrations and correlated them with substrate processing and protein structural changes. Chondroitin sulfate (CS) and heparan sulfate (HS) bound to K5 lyase with a Kd of 0.5 microM whereas heparin exhibited a Kd=1.1 microM. The natural substrate K5 polysaccharide displayed a similar apparent affinity as CS and HS but was the only ligand of the enzyme which induced a large structural rearrangement of the protein as detected by far-UV CD spectroscopy. Since significant enzymatic degradation was only found for the K5 polysaccharide peaking at 44 degrees C, but binding was also detected for heparin, we propose that the K5 lyase is able to discriminate between specific (acetylated/non-sulfated) and unspecific (acetylated/sulfated) ligands by its heparin binding motif in the C-terminus. This is proposed to be the origin for the enzyme's residual HS degrading activity.  相似文献   

12.
P Midoux  A C Roche  M Monsigny 《Cytometry》1987,8(3):327-334
The fluorescence properties of the fluorescein residues bound to a protein are used to analyze by flow cytometry the neoglycoproteins' endocytosis mediated by membrane lectins of Lewis lung carcinoma cells (3LL cells). The quantum yield of fluorescein bound to a protein is dependent on the number of fluorophore molecules bound to a protein molecule and the pH of the environmental medium. The mean fluorescence intensity of a fluorescein molecule bound to a protein decreases when the number of fluorescein residues per protein molecule increases. However, after proteolytic digestion, the mean fluorescence intensity of a fluorescein molecule is constant and equal to that of free fluorescein. The binding of fluorescein-labeled alpha-glucosylated serum albumin to 3LL cells at 4 degrees C can easily be determined by flow cytometry because under these conditions the environmental pH is neutral, and the neoglycoprotein is not degraded. When the cells are incubated at 37 degrees C in the presence of a fluorescein-labeled neoglycoprotein, the fluorescence intensity of a cell is low because of the low pH of endosomes and lysosomes but is increased upon a postincubation at 4 degrees C in the presence of monensin, a proton/sodium ionophore. The extent of the proteolytic digestion of an endocytosed neoglycoprotein can be assessed by comparing, upon a monensin postincubation at 4 degrees C, the high cell-associated fluorescence of cells incubated in the absence of leupeptin (an inhibitor of lysosomal proteases) and the relatively low fluorescence intensity of cells incubated in the presence of leupeptin.  相似文献   

13.
The mechanism of the heparin-promoted reaction of thrombin with antithrombin III was investigated by using covalent complexes of antithrombin III with either high-affinity heparin (Mr = 15,000) or heparin fragments having an average of 16 and 12 monosaccharide units (Mr = 4,300 and 3,200). The complexes inhibit thrombin in the manner of active site-directed, irreversible inhibitors: (Formula: see text) That is, the inhibition rate of the enzyme is saturable with respect to concentration of complexes. The values determined for Ki = (k-1 + k2)/k1 are 7 nM, 100 nM, and 6 microM when the Mr of the heparin moieties are 15,000, 4,300, 3,200, respectively, whereas k2 (2 S-1) is independent of the heparin chain length. The bimolecular rate constant k2/Ki for intact heparin is 3 X 10(8) M-1 S-1 and the corresponding second order rate constant k1 is 6.7 X 10(8) M-1 S-1, a value greater than that expected for a diffusion-controlled bimolecular reaction. The bimolecular rate constants for the complexes with heparin of Mr = 4,300 and 3,200 are, respectively, 2 X 10(7) M-1 S-1 and 3 X 10(5) M-1 S-1. Active site-blocked thrombin is an antagonist of covalent antithrombin III-heparin complexes: the effect is monophasic and half-maximum at 4 nM of antagonist against the complex with intact heparin, whereas the effect is weaker against complexes with heparin fragments and not monophasic. We conclude that virtually all of the activity of high affinity, high molecular weight heparin depends on binding both thrombin and antithrombin III to heparin, and that the exceptionally high activity of heparin results in part from the capacity of thrombin bound nonspecifically to heparin to diffuse in the dimension of the heparin chain towards bound antithrombin III. Increasing the chain length of heparin results in an increased reaction rate because of a higher probability of interaction between thrombin and heparin in solution.  相似文献   

14.
The fluorescence properties of fluorescein bound to protein are used to quantitate by flow cytofluorometry the degradation of fluorescein-labeled alpha-glucosylated serum albumin (fluorescein-labeled neoglycoprotein) after endocytosis by the membrane lectin of Lewis lung carcinoma cells (3LL cells). The quantum yield of fluorescein bound to a protein decreases when the number of fluorescein residues per protein molecule increases; however, after proteolytic digestion the mean fluorescence intensity of a fluorescein molecule is constant and equal to that of free fluorescein. The extent of the degradation of the endocytosed neoglycoprotein was determined with a flow cytofluorometer by using two neoglycoproteins containing either a small or a large number of fluorescein residues per neoglycoprotein molecule. At 4 degrees C, 3LL cells bind 750,000 molecules of fluorescein-labeled alpha-glucosylated serum albumin with an apparent binding constant of 2 X 10(6) 1 X mole-1. At 37 degrees C, after 4 hr incubation 2.2 X 10(6) molecules of fluorescent alpha-glucosylated serum albumin were cell-associated, and of these at least one third were degraded.  相似文献   

15.
We have studied at a molecular level the interaction of heparins on bothropstoxin-I (BthTx-I), a phospholipase A2 toxin. The protein was monitored using gel filtration chromatography, dynamic light scattering (DLS), circular dichroism (CD), attenuated total reflectance Fourier transform infrared (ATR-FTIR) and intrinsic tryptophan fluorescence emission (ITFE) spectroscopy. The elution profile of the protein presents a displacement of the protein peak to larger complexes when interacting with higher concentration of heparin. The DLS results shows two Rh at a molar ratio of 1, one to the distribution of the protein and the second for the action of heparin on BthTx-I structures, and a large distribution with the increase of protein. The interaction is accompanied by significant changes in the CD spectra, showing two common features: a decrease in signal at 208 nm (3 and 6 kDa heparins) and an isodichroic point near 226 nm (3 kDa heparin). FTIR spectra indicate that only a few amino acid residues are involved in this interaction. Alterations in the ITFE by binding heparins suggest that the initial binding occurs on the ventral face of BthTx-I. Together, these results add an experimental and structural basis on the action mechanism of the heparins over the phospholipases A2 and provide a molecular model to elucidate the interaction of the enzyme-heparin complex at a molecular level.  相似文献   

16.
At initial stages in the Rhizobium legume symbiosis, most nodulation genes are controlled by NodD protein and plant inducers. Some genetic studies and other reports have suggested that NodD may be activated by its direct interaction with plant inducers. However, there has been no molecular evidence of such an inducing interaction. In this paper, we used fluorescence resonance energy transfer technique to see whether such an interaction exists between NodD and its activator, naringenin, in vitro. The tetracysteine motif (Cys-Cys-Pro-Gly-Cys-Cys) was genetically inserted into NodD to label NodD with 4',5'-bis(1,3,2-dithioarsolan-2-yl) fluorescein (FlAsH). Naringenin was labeled with fluorescein by chemical linking. In the fluorescence resonance energy transfer experiments in vitro, the fluorescence intensity of one acceptor, NodD(90R6)-FlAsH, increased by 13%. This suggests that NodD may directly interact with inducer naringenin in vitro and that the reaction centre is likely near hinge region 1 of NodD.  相似文献   

17.
Heparan sulfate is a polysaccharide belonging to the glycaminoglycan family. It interacts with numerous proteins of the extracellular matrix, in particular cellular growth factors. The number of experimental protein-heparin sulfate complexes obtained by crystallography or nuclear magnetic resonance is limited. Alternatively, computational approaches can be employed. Generally, they restrain the conformation of the glycosidic rings and linkages in order to reduce the complexity of the problem. Modeling the interaction between protein and heparan sulfate is indeed challenging because of the large size of the fragment needed for a strong binding, the flexibility brought by the glycosidic rings and linkages and the high density of negative charges. We propose a two-step method based on molecular docking and molecular dynamics simulation. Molecular docking allows exploring the positioning of a rigid heparin sulfate fragment on the protein surface. Molecular dynamics refine selected docking models by explicitly representing solvent molecules and not restraining the polysaccharide backbone. The interaction of a hexamer of heparin sulfate was studied in interaction with fibroblast growth factor 2 and stromal cell-derived factor 1α. This approach shed light on the plasticity of the growth factors interacting with heparan sulfate. This approach can be extended to the study of other protein/glycosaminoglycan complexes.  相似文献   

18.
The effect of heparin fractions of various Mr, with high affinity for antithrombin III, on the kinetics of the reaction between factor Xa and antithrombin III have been studied using purified human proteins. Each of the heparin fractions, which varied between pentasaccharide and Mr 32,000, accelerated the inhibition of factor Xa although an increasing rate of inhibition was observed with increasing Mr. The chemically synthesized pentasaccharide preparation (Mr 1714) gave a maximum inhibition rate constant of 1.2 X 10(7) M-1 X min-1, compared with 6.3 X 10(4) M-1 X min-1 in the absence of heparin, and this rose progressively to 4.2 X 10(8) M-1 X min-1 with the two fractions of highest Mr (22,500 and 32,000). The 35-fold difference in inhibition rates observed with the high-affinity fractions was virtually abolished by the presence of 0.3 M-NaCl. The disparity in these rates of inhibition was shown to be due to a change in the Km for factor Xa when a two-substrate model of heparin catalysis was used. The Km for factor Xa rose from 28 nM for the fraction of Mr 32,000 to 770 nM for the pentasaccharide, whilst 0.3 M-NaCl also caused an increase in Km with the high-Mr fraction. These data suggest that the increased rates of inhibition observed with heparins of higher Mr may be due to an involvement of heparin binding to factor Xa as well as to antithrombin III.  相似文献   

19.
Limited proteolysis of high molecular weight kininogen by kallikrein resulted in the generation of an inactive heavy chain of Mr = 64,000 and active light chains of Mr = 64,000 and 51,000 when analyzed by sodium dodecyl sulfate (SDS)-gel electrophoresis under reducing conditions. Starting with kininogen from outdated plasma, a light chain with an apparent molecular weight of 51,000 on 7.5% SDS gels was purified and characterized. Molecular weights of 28,900 +/- 1,100 and 30,500 +/- 1,600 were obtained by gel filtration of the reduced and alkylated protein in 6 M guanidine HCl and equilibrium sedimentation under nondenaturing conditions in the air-driven ultracentrifuge, respectively. The light chain stained positively with periodic acid-Schiff reagent on SDS gels indicating that covalently attached carbohydrate may be responsible for the anomalously high molecular weight estimated by SDS-gel electrophoresis. A single light chain thiol group reacted with 5,5'-dithiobis-(2-nitrobenzoic acid) (DTNB) in the presence and absence of 6 M guanidine HCl. Specific fluorescent labeling of the thiol group with 5-(iodoacetamido)fluorescein (IAF) occurred without loss of clotting activity. Addition of purified human plasma prekallikrein to the IAF-light chain resulted in a maximum increase in fluorescence anisotropy of 0.041 +/- 0.001 and no change in the fluorescence intensity. Fluorescence anisotropy measurements of the equilibrium binding of prekallikrein to the IAF-light chain yielded an average Kd of 17.3 +/- 2.5 nM and stoichiometry of 1.07 +/- 0.07 mol of prekallikrein/mol of IAF-light chain. Measurements of the interaction of prekallikrein with iodoacetamide-alkylated light chain using the IAF-light chain as a probe gave an average Kd of 16 +/- 4 nM and stoichiometry of 1.0 +/- 0.2 indicating indistinguishable affinities for prekallikrein.  相似文献   

20.
Coated vesicles isolated from bovine brain contained a protein kinase(s) which phosphorylated phosvitin and an endogenous protein with a molecular weight (Mr) of 48,000. A clathrin light chain (Mr 33,000), a constituent of the coat structure of the coated vesicles, was also phosphorylated when histone was added to the incubation medium. The clathrin light chain was phosphorylated with GTP as well as ATP as the phosphoryl donor. The phosphorylation reaction was inhibited by heparin. An additional 1.35 mol of PO4/mol was incorporated into the clathrin light chain which had contained approximately 1.5 mol of PO4/mol when the coated vesicles were incubated with ATP, Mg2+, and histone. Phosphoamino acid determination revealed the presence of 32P-phosphorylated threonine and serine in phosvitin, threonine in the endogenous protein (Mr 48,000) and serine in the clathrin light chain (Mr 33,000).  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