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 46 毫秒
1.
This study characterizes the physical-chemical interactions of heparin with human plasma low-density lipoproteins (LDL). A high reactive heparin (HRH) specific for the surface determinants of LDL was isolated by chromatography of commercial bovine lung heparin on LDL immobilized to AffiGel-10. HRH was derivatized with fluoresceinamine and repurified by affinity chromatography, and its interaction with LDL in solution was monitored by steady-state fluorescence polarization. Binding of LDL to fluoresceinamine-labeled HRH (FL . HRH) was saturable, reversible, and specific; HRH stoichiometrically displaced FL . HRH from the soluble complex, and acetylation of lysine residues on LDL blocked heparin binding. Titration of FL.HRH with excess LDL yielded soluble complexes with two LDL molecules per heparin chain (Mr 13,000) characterized by an apparent Kd of 1 microM. Titration of LDL with excess HRH resulted in two classes of heparin binding with two and five heparin molecules bound per LDL and apparent Kd values of 1 and 10 microM, respectively. At physiological pH and ionic strength, the soluble HRH-LDL complexes were maximally precipitated with 20-50 mM Ca2+. Insoluble complexes contained 2-10 HRH molecules per LDL with the final product stoichiometry dependent on the ratio of the reactants. The affinity of HRH for LDL in the insoluble complexes was estimated between 1 and 10 microM. Insoluble LDL-heparin complexes were readily dissociated with 1.0 M NaCl, and their formation was prevented by acetylation of the lysine residues on LDL.  相似文献   

2.
A ligand-blotting procedure which allows detection of heparin-binding proteins is described. Crude commercial heparin was fractionated by chromatography on a column of human plasma low-density lipoproteins immobilized to Sepharose CL-4B. Chromatography yielded an unbound and a bound fraction of heparin, designated URH and HRH, respectively. The HRH fraction was reacted with the N-hydroxysuccinimidyl ester of 3-(p-hydroxyphenyl)propionic acid and then labeled with 125I. Proteins were separated by 3-20% pore-gradient gel electrophoresis, transferred to nitrocellulose, and then assayed for their ability to bind 125I-labeled HRH. Human plasma apolipoproteins B-100, B-48, and E of chylomicrons, very low-density lipoproteins, and low-density lipoproteins bound the 125I-labeled HRH; the radiolabeled heparin did not bind to serum albumin, ferritin, catalase, and lactate dehydrogenase. The ligand-blotting procedure should facilitate the purification of heparin-binding domains from these proteins and, moreover, may be applicable to the investigation of heparin-protein interactions in general.  相似文献   

3.
Human plasma low-density lipoproteins (LDL) were incubated with 10 microM probucol for 1 h at 37 degrees C. Probucol incorporation into the LDL was complete as judged by filtration through a 0.2-micron filter, ultracentrifugation, and gel filtration. LDL with and without probucol were incubated for up to 24 h with 5 microM Cu2+ at 37 degrees C. Copper oxidation increased the content of random structure in the LDL protein from 30% to 36% at the expense of beta-structure (which decreased from 22% to 16%) without a change in alpha-helical content as measured by circular dichroism spectroscopy. This loss of beta-structure was prevented by the presence of probucol in the LDL during the copper incubation. Probucol reduced the rate of increase of fluorescence during copper oxidation at 37 degrees C. After 6 h, the fluorescence intensity at 360-nm excitation and 430-nm emission was 30% less in probucol-containing samples. Probucol had no effect on the circular dichroic spectrum of LDL and only minimal effects (less than 5%) on the fluorescence emission spectrum at wavelengths below 500 nm. Two fluorescence peaks, with emission at 420 nm and excitation at 340 and 360 nm, are resolved in three-dimensional fluorescence spectra of oxidized LDL. Probucol reduces the intensity of both peaks equally. The binding of a highly reactive heparin (HRH) fraction to LDL was measured by titration of LDL with HRH in the presence of fluoresceinamine-labeled HRH. The decrease in fluorescence anisotropy of the labeled HRH is proportional to the concentration of bound HRH.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
Heparin was fractionated on an affinity column of bovine milk lipoprotein lipase (LpL) immobilized to Affi-Gel-15. The bound heparin, designated high-reactive heparin (HRH), enhanced LpL activity, presumably by stabilizing the enzyme against denaturation. The unbound heparin fraction had no observable effect on the initial rate of enzyme activity. However, at longer times of incubation there was inhibition of LpL activity. LpL-specific HRH also showed a high, Ca2+-dependent precipitating activity towards human plasma low density lipoproteins (LDL). Since LpL and LDL both bind to heparin-like molecules at the surface of the arterial wall, we suggest that their similar heparin-binding specificity may have physiological consequences as it relates to the development of atherosclerosis.

Heparin binding Lipoprotein lipase LDL Apolipoprotein Lipolysis  相似文献   


5.
The binding of [3H]heparin to human plasma lipoproteins was measured using a gel filtration assay on columns of Ultrogel AcA 54. [3H]Heparin formed a soluble complex with low density lipoprotein (LDL) as evidenced by the appearance of a new radioactive peak emerging at the void volume where the lipoproteins elute. Free heparin on the other hand was retarded on this column and eluted at a later volume. Heparin binding to LDL could also be demonstrated on columns of Sepharose 4B, in which case two included peaks of 3H were observed to elute in the area of LDL and of heparin. [3H]Heparin did not bind to either high or very low density lipoproteins as determined by the gel filtration assay. The binding of the [3H]heparin to LDL was proportional to both the concentration of LDL and of heparin and both showed saturation kinetics. Cations were not necessary for binding, nor was binding inhibited by EDTA. LDL showed a marked specificity for heparin. Thus, the binding of [3H]heparin to LDL was strongly inhibited by the addition of unlabeled heparin, while other glycosaminoglycans such as chondroitin sulfate, heparan sulfate, keratan sulfate, and dermatan sulfate were not effective inhibitors except at very high concentrations. Salts, especially K2HPO4 and (NH4)2SO4, also inhibited binding when added at concentrations of 10 mm or higher suggesting an ionic interaction between heparin and LDL. The pH optimum for binding was between 7.5 and 8.5 but binding fell off markedly above pH 9.0. The [3H]heparin was heterogeneous and could be separated into four fractions on columns of Sephadex G-75. When these fractions were tested for binding to LDL, only the high molecular weight fraction bound to any significant extent. LDL was treated with reagents used to selectively modify basic amino acid residues, and the effect of these treatments on heparin binding was examined. Thus, ethoxyformic anhydride was used for histidine modification, acetic anhydride and succinic anhydride for lysines and cyclohexanedione for arginine residues. In each case there was a significant loss in heparin binding suggesting that various basic amino acids are involved in binding and/or that basic amino acids are necessary to maintain the proper conformation of LDL.  相似文献   

6.
Three distinctive heparin-binding sites were observed in type IV collagen by the use of rotary shadowing: in the NC1 domain and at distances 100 and 300 nm from the NC1 domain. Scatchard analysis indicated different affinities for these sites. Electron microscopic analysis of heparin-type IV collagen interaction with increasing salt concentrations showed the different affinities to be NC1 greater than 100 nm greater than 300 nm. The NC1 domain bound specifically to chondroitin/dermatan sulfate side chains as well. This binding was observed at the electron microscope and in solid-phase binding assays (where chondroitin sulfate could compete for the binding of [3H]heparin to NC1-coated substrata). The triple helix-rich, rod-like domain of type IV collagen did not bind to chondroitin/dermatan sulfate side chains. In solid-phase binding assays only heparin could compete for the binding of [3H]heparin to this domain. In order to more precisely map potential heparin-binding sites in type IV collagen, we chemically synthesized 17 arginine- and lysine-containing peptides from the alpha 1(IV) and alpha 2(IV) chains. Three peptides from the known sequence of the alpha 1(IV) and alpha 2(IV) chains were shown to specifically bind heparin: peptide Hep-I (TAGSCLRKFSTM), from the alpha 1(NC1) chain, peptide Hep-II (LAGSCLARFSTM), a peptide corresponding to the same sequence in peptide Hep-I from the alpha 2 (NC1) chain, and peptide Hep-III (GEFYFDLRLKGDK) which contained an interruption of the triple helical sequence of the alpha 1(IV) chain at about 300 nm from the NC1 domain, were demonstrated to bind heparin in solid-phase binding assays and compete for the binding of [3H]heparin to type IV collagen-coated substrata. Therefore, each of these peptides may represent a potential heparin-binding site in type IV collagen. The mapping of the binding of heparin or related structures, such as heparan sulfate proteoglycan, to specific sequences of type IV collagen could help the understanding of several structural and functional properties of this basement membrane protein as well as interactions with other basement membrane and/or cell surface-associated macromolecules.  相似文献   

7.
Binding and endocytosis of heparin by human endothelial cells in culture   总被引:8,自引:0,他引:8  
Binding of heparin and low molecular weight heparin fragments (CY 222, Mr range 1500-8000) to human vascular endothelial cells was studied. Primary culture of human umbilical vein endothelial cells and either 125I or 3H-labeled heparin or [125I]CY 222 were used. Slow, saturable and specific binding was found. No other tested glycosaminoglycan, excepting a highly sulfated heparan fraction, was able to compete for heparin binding. Two groups of binding sites for [3H]heparin could be distinguished: one with high affinity (Kd = 0.12 microM) and another with lower affinity (Kd = 1.37 microM) and a relative large capacity of binding (1.16 X 10(7) molecules/cell) was calculated. The Kd for unlabeled heparin, as calculated from competition experiments, was 0.23 microM. Much lower affinity was calculated for unlabeled low molecular weight heparin fragments CY 222 (Kd = 4.3 microM) from competition experiments with [125I]CY 222. The binding reversibility was only partial for unfractionated heparin. Even by chasing with unlabeled compound, a fraction of 25-30% was not dissociable from endothelial cells. This fraction was much lower if incubation was carried out at 4 degrees C. The addition of basic proteins (histones) to the incubation medium greatly enhanced the undissociable binding at 37 degrees C, but not at 4 degrees C. The undissociable fraction of heparin was not available to degradation by purified microbial heparinase. These results suggest that a fraction of bound heparin is internalized by the vascular endothelium.  相似文献   

8.
Binding of radiolabeled (125I) heparin and its low molecular weight fraction CY 222 (Choay) to human and porcine cultured endothelium was investigated. The binding was measured over a wide range of heparin or CY 222 concentration in culture medium, from less than 20-time up to more than 30-times of the therapeutic heparin level. A relatively small fraction (less than 1%) of tested products was bound to endothelium. The process of binding was temperature-independent. A comparable number of endothelial binding sites (approx. 10(12)/cm2) for both 125I-heparin and 125I-CY 222 was calculated. About 40% totally bound 125I-heparin and 30% of 125I-CY 222 was found in extracellular matrix of cultured endothelium. The endothelium exhibited a 2.4-times lower affinity for 125I-CY 222 (Kd = 5.59 +/- 1.77 microM) than that for 125I-heparin (Kd = 2.35 +/- 0.78 microM). A similar affinity of human (venous) and porcine (aortic) endothelium for 125I-heparin was demonstrated. Cultured endothelium exhibited the same affinity for unlabeled heparin as for 125I-heparin. The endothelium depleted of sialic acid residues bound 1.5-times more of 125I-CY 222 than the control endothelium in culture.  相似文献   

9.
The binding of collagens and fragments of type I collagen to heparin was studied by gel electrophoresis and affinity chromatography. Samples bound in 150 mM NaCl/10 mM Hepes (pH6.5) were eluted with 2 M NaCl, 6 M urea, or a linear gradient of 0.15–1.0 M NaCl. The triple-helical conformation was shown to be essential for binding. The vertebrate collagenase-generated C-terminal fragment, TCB was shown to have greater binding affinity for heparin than the N-terminal TCA fragment. Both type II collagen and the NC1 domain of type IV collagen bound to heparin, whereas pepsin-solubilized tetrameric type IV failed to bind.  相似文献   

10.
The specificity of endothelial binding sites for heparin was investigated with heparin fractions and fragments differing in their Mr, charge density and affinity for antithrombin III, as well as with heparinoids and other anionic polyelectrolytes (polystyrene sulphonates). The affinity for endothelial cells was estimated by determining I50 values in competition experiments with 125I-heparin. We found that affinity for endothelial cells increases as a function of Mr and charge density (degree of sulphation). Binding sites are not specific receptors for heparin. Other anionic polyelectrolytes, such as pentosan polysulphates and polystyrene sulphonates, competed with heparin for binding to endothelial cells. Fractions of standard heparin with high affinity for antithrombin III also had greater affinity for endothelium. However, these two properties of heparin (affinity for antithrombin III and affinity for endothelial cells) could be dissociated. Oversulphated heparins and oversulphated low-Mr heparin fragments had lower anticoagulant activity and higher affinity for endothelial cells than did their parent compounds. Synthetic pentasaccharides, bearing the minimal sequence for binding to antithrombin III, did not bind to endothelial cells. Binding to endothelial cells involved partial neutralization of heparin. Bound heparin exhibited only 5% and 7% of antifactor IIa and antifactor Xa specific activity, respectively. In the presence of 200 nM-antithrombin III, and in the absence of free heparin, a limited fraction (approx. 30%) of bound heparin was displaced from endothelial cells during a 1 h incubation period. These data suggested that a fraction of surface-bound heparin could represent a pool of anticoagulant.  相似文献   

11.
Preparations of low molecular weight porcine heparin with an average specific anticoagulant activity of 94 units/mg were fractionated into "active" and "relatively inactive" forms of the mucopolysaccharide of approximately 6000 daltons each. The active fraction was further subdivided into various species with descending but significant affinities for the protease inhibitor as well as decreasing but substantial anticoagulatn potencies. "Highly active" heparin (approximately 8% of the low molecular weight pool) possesses a specific anticoagulant activity of 350 +/- 10 units/mg. The relatively inactive fraction (67% of the low molecular weight pool) exhibits a specific anticoagulant activity of 4 +/- 1 units/mg. The binding of highly active heparin to antithrombin is accurately described by a single-site binding model with a KHep-ATDISS of approximately 1 X 10(-7) M. Variations in this binding parameter secondary to changes in environmental variables indicate that charge-charge interactions as well as an increase in entropy are critical to the formation of the highly active heparin-antithrombin complex. The interaction of relatively inactive heparin with the protease inhibitor is characterized by an apparent KHep-ATDISS of 1 X 10(-4) M. In large measure, this is due to small amounts of residual active mucopolysaccharide (0.5%). The ability of the highly active heparin to accelerate the thrombin-antithrombin interaction was also examined. We were able to demonstrate that the mucopolysaccharide acts as a catalyst in this process and is able to initiate multiple rounds of enzyme-inhibitor complex formation. The rate of enzyme neutralization is increased to a maximum of 2300-fold as the concentration of heparin is raised until the inhibitor is saturated with mucopolysaccharide. Further increases in heparin concentration result in a reduction in the speed of enzyme neutralization. This appears to be due to the formation of thrombin-heparin complexes. A mathematical model is given which provides a relationship between the initial velocity of enzyme neutralization and reactant concentrations.  相似文献   

12.
Apolipoprotein B-100 (apoB-100) is the major protein constituent of human plasma low-density lipoproteins (LDL). On the basis of its amino acid sequence [Chen, S.-H., Yang, C.-Y., Chen, P.-F., Setzer, D., Tanimura, M., Li, W.-H., Gotto, A. M., Jr., & Chan, L. (1986) J. Biol. Chem. 261, 12918-12921], apo B-100 is one of the largest monomeric proteins known with a calculated molecular weight of 512937. Heparin binds to the LDL surface by interacting with positively charged amino acid residues of apoB-100, forming soluble complexes in the absence of divalent metals and insoluble complexes in their presence. The purpose of this study was to isolate and characterize the heparin-binding domain(s) of apoB-100. Human plasma LDL were fragmented with cyanogen bromide (CNBr). After delipidation and reduction-carboxymethylation, the CNBr peptides were fractionated by sequential chromatography on DEAE-Sephacel, Mono S, and high reactive heparin (HRH) AffiGel-10; HRH was purified by chromatography of crude bovine lung heparin on LDL AffiGel-10. Heparin-binding peptides were further purified by reverse-phase high-performance liquid chromatography. Heparin-binding activity was monitored by a dot-blot assay with 125I-HRH. The amino-terminal sequences of four CNBr heparin-binding peptides (CNBr-I-IV) were determined. CNBr-I-IV correspond to residues 2016-2151, 3109-3240, 3308-3394, and 3570-3719, respectively, of the amino acid sequence of apoB-100. Each CNBr peptide contains a domain(s) of basic amino acid residues which we suggest accounts for their heparin-binding activity.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
The C-terminal domain of lipoprotein lipase (LPL) is involved in several important interactions. To assess its contribution to the binding ability of full-length LPL we have determined kinetic constants using biosensor technique. The affinity of the C-terminal domain for heparin was about 500-fold lower than that of full-length LPL (K(d) = 1.3 microM compared to 3.1 nM). Replacement of Lys403, Arg405 and Lys407 by Ala abolished the heparin affinity, whereas replacement of Arg420 and Lys422 had little effect. The C-terminal domain increased binding of chylomicrons and VLDL to immobilized heparin relatively well, but was less than 10% efficient in binding of LDL compared to full-length LPL. Deletion of residues 390-393 (WSDW) did not change the affinity to heparin and only slightly decreased the affinity to lipoproteins. We conclude that the C-terminal folding domain contributes only moderately to the heparin affinity of full-length LPL, whereas the domain appears important for tethering triglyceride-rich lipoproteins to heparin-bound LPL.  相似文献   

14.
The effect of the polyanion heparin on the trehalose phosphate synthetase of Mycobacterium smegmatis had been studied. In the presence of heparin (0.5 mg/ml), the synthetase shows greatly increased stability when heated at 50 °C for various periods of time as compared to the enzyme in the absence of heparin. Heparin also prevents digestion of the enzyme by trypsin. In the absence of heparin, the synthetase is retained on a Sephadex G-200 column and elutes in an area suggesting a molecular weight of about 40,000–50,000. However, when heparin (0.5 mg/ml) is mixed with the enzyme, the synthetase is excluded from the Sephadex G-200 column and elutes in an area suggesting a molecular weight of greater than 450,000. The trehalose phosphate synthetase was purified by binding it to a column of heparin covalently attached to Sepharose 4B. The synthetase was eluted from this column with a linear gradient of heparin. This enzyme fraction which contained bound heparin showed greatly increased stability at 50 °C, and eluted from the Sephadex G-200 column in an area suggesting a molecular weight of greater than 450,000. These results indicate that heparin, and presumably other polyanions, stabilizes the synthetase to adverse conditions and also causes an association of the enzyme to high molecular weight forms.The synthetase, when bound to the heparin-Sepharose gel, still retained good enzymatic activity. This immobilized enzyme was active with various glucose sugar nucleotides (ADP-glucose, GDP-glucose, UDP-glucose, TDP-glucose) and did not require additional polyanion. The product formed from each of these sugar nucleotides was shown to be trehalose phosphate by a variety of chemical and enzymatic procedures.  相似文献   

15.
Low molecular weight heparin of low-anticoagulant activity and high molecular weight heparin of correspondingly high activity were prepared by chromatography on protamine-Sepharose; preparations subjected to limited N-desulfation (5–10% free amino groups) by solvolysis were labeled with 5-dimethylaminonaphthalene-1-sulfonyl chloride (dansyl chloride) or rhodamine B isothiocyanate (RITC). The fluorescent heparins retained approximately 50% of the original anticoagulant activities. Dansyl-heparin on binding to antithrombin III (ATIII) exhibited a 2.5-fold enhancement of dansyl fluorescence intensity. This effect could be prevented by excess unlabeled heparin. A 7900 molecular weight dansyl-heparin preparation bound to ATIII with a stoichiometry of close to 2:1 and with an apparent association constant for binding (Ka) of 4.9 × 105, m?1, whereas a 21,600 molecular weight fraction bound at 0.7:1 with the protein and with an apparent Ka = 7.9 × 105, m?1. When ATIII reacted with a mixture of low molecular weight dansyl-heparin and low molecular weight RITC-heparin, there was enhancement of RITC fluorescence emission when excited at the dansyl excitation maximum; this effect was not observed when either of the labeled heparin species was prepared from high molecular weight material. The results are consistent with the proposal that a single molecule of high molecular weight, high-activity heparin occupies two sites when it binds to ATIII, whereas low molecular weight, low-activity heparin binds to the two sites separately.  相似文献   

16.
Heparin binding to human low density lipoproteins (LDL) and the effect of heparin on the ability of LDL to bind to the LDL receptor has been investigated. Emphasis has been made on the physiological conditions of temperature, pH and the ionic strength. Intrinsic fluorescence spectroscopy of LDL has been applied to follow heparin binding. Fluorescence anisotropy has been measured to describe the changes in apoB and dansyl-heparin dynamics upon binding. Eu3+-labeled LDL binding to the intact LDL receptor has been monitored by time-resolved fluorescence spectroscopy technique. We have found that heparin binds to LDL under the physiological conditions, probably by Van der Waals interactions and hydrogen bonding. Temperature seems to be the most important factor influencing the interaction. Furthermore, the presence of heparin inhibits LDL binding to the intact LDL receptor that might have consequences on the cholesterol metabolism in vivo.  相似文献   

17.
In this study, the interaction of human serum low-density lipoprotein (LDL) with heparin immobilized on Sepharose was reinvestigated. Binding of isolated LDL (stabilized with human serum albumin (HSA] was compared with that of LDL in full serum. (1) Binding of isolated LDL was slightly decreased by CaCl2 and was not affected by MgCl2. In contrast, with full serum LDL binding was increased by these divalent cations. (2) In both situations, binding of LDL was saturable, but the maximum degree of binding that could be reached was much higher with isolated LDL than with LDL in full serum. This could be ascribed to an inhibitory action of a factor found in the d greater than 1.24 fraction of serum. (3) The effect of this factor was diminished in the presence of CaCl2 or MgCl2, which suggests that the stimulation of LDL binding by these cations in full serum is due to suppression of the inhibitory activity of this factor. (4) The inhibitory factor in the d greater than 1.24 fraction can be partially purified by absorption to heparin-Sepharose, followed by elution with 6 M guanidine chloride. The resulting preparation had a 30- to 50-fold higher specific activity. Attempts to purify the factor further resulted in loss of activity. (5) The activity is decreased upon treatment with trypsin and also upon acetylation or reduction with dithiothreitol, indicating that free amino groups and S-S bridges are essential.  相似文献   

18.
Heparin was introduced into the stomach or duodenum of mice separately in doses of ca. 250 mg/kg. A slight anticoagulant effect in the systemic circulation was detected in whole blood clotting times and factor X inhibition. In contrast to most drugs, more heparin was absorbed from the stomach than from the intestine. Suppressing ionization of heparin by simultaneous administration of acid resulted in improved absorption of heparin from the small intestine. Heparin was separated with ethanol into five molecular weight fraction: I, 17 999; II, 13 i99; III, 10800, IV, 8 700; and V, 6 700. Each was introduced into the duodenum of mice with citric acid. The maximum hypocoagulability was produced with fraction IV. When administered in distilled water instead of in citric acid, this heparin fraction did not produce an anticoagulant effect. These studies demonstrated that improvement of heparin absorption from the gastrointestinal tract can be obtained by the combination of suppressing ionization and selecting molecular size.  相似文献   

19.
Binding of heparin to human high molecular weight kininogen   总被引:1,自引:0,他引:1  
The binding of heparin to high molecular weight kininogen (H-kininogen) was analyzed by the effect of kininogen in decreasing the heparin-induced enhancement of the rate of inactivation of thrombin by antithrombin. The conditions were arranged so that the heparin-catalyzed antithrombin-thrombin reaction, monitored in the presence of the reversible thrombin inhibitor p-aminobenzamidine, followed pseudo-first-order kinetics and the observed rate constant (kappa obsd) varied linearly with the heparin concentration. In the absence of metal ions, H-kininogen minimally affected kappa obsd, measured at a constant concentration of heparin with high affinity for antithrombin (30 nM), at I = 0.15, pH 7.4 and 25 degrees C. However, at a saturating concentration of Zn2+ (10 microM), kappa obsd was reduced to 50% at approximately 20 nM H-kininogen and to that of the uncatalyzed reaction at greater than or equal to approximately 0.2 microM H-kininogen. Conversely, at a saturating concentration of H-kininogen (0.5 microM), kappa obsd was decreased to 50% at approximately 0.6 microM Zn2+ and to the kappa obsd of the uncatalyzed reaction at greater than or equal to 10 microM Zn2+. Other metal ions were effective in the order Zn2+ approximately Ni2+ greater than Cu2+ approximately Co2+ approximately Cd2+. The single-chain and two-chain forms of H-kininogen and the H-kininogen light chain reduced the heparin enhancement in the presence of Zn2+ to the same extent, whereas low molecular weight kininogen had no influence.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
B Lages  S S Stivala 《Biopolymers》1973,12(5):961-974
The binding of copper ions to fibrinogen was studied by the equilibrium dialysis technique in neutral Tris buffer. The presence of copper causes precipitation of fibrinogen-copper complexes, the amount of which varies with the copper ion concentration. Solutions of 96% clottable fibrinogen (whole fibrinogen) displayed a maximum binding capacity about four times greater than that of fibrinogen solutions from which the cold-insoluble precipitate had been removed (cold-soluble fibrinogen). Binding in both systems apparently involves two classes of sites in fibrinogen, and the class of lower affinity is associated with cooperative interactions with copper. The copper concentration at which this cooperative uptake occurs is identical to the concentration at which the amount of precipitated material increases sharply and also to the concentration at which a sharp decrease is observed in the sedimentation coefficient of soluble fibrinogen, suggesting some relationship between copper binding, solubility, and solution properties. The presence of heparin markedly affects the sedimentation coefficient of fibrinogen in the presence of copper ions, although showing a lesser effect in the absence of the metal. The sedimentation coefficient of fibrinogen is increased in the presence of heparin and copper ion, compared to the value of fibrinogen-copper systems without heparin, and this effect is enhanced by changing the fibrinogen:heparin molar ratio to larger values. The precipitation of fibrinogen from solution, apparently without a coincident removal of heparin, also increases with increasing copper ion concentration and fibrinogen:heparin molar ratio. The possible significance of these effects in terms of heparin anti-coagulant activity is discussed.  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