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 421 毫秒
1.
Binding of Phenol Red to human serum albumin at pH 7.0 was studied by ultrafiltration (n1 = 1, K1 = 3.9 X 1-(4) M-1, n2 = 5, K2 = 9.6 X 10(2) M-1). The presence of 1 mol of octanoate or decanoate per mol of albumin caused a decrease in dye binding (dye/protein molar ratio 1:1), which, in contrast with additional fatty acid, was very pronounced: 1-8 mol of palmitate or stearate resulted in a small, and apparently linear, displacement of Phenol Red. The displacement effect of 1-5 mol of oleate, linoleate or linolenate per mol of albumin was comparable with that of the equimolar concentrations of palmitate or stearate. A higher molar ratios the unsaturated acids caused a drastic decrease in dye binding. The different Phenol Red-displacement effects of low molar ratios of medium-chain and long-chain fatty acids indicate that these acids have different high-affinity binding sites. In accordance with this proposal, low concentrations of stearate had only a small effect on the Phenol Red-displacement effect of octanoate. Phenol Red-binding curves in the presence of 1 mol of octanoate, 8 mol of stearate and 6 or 7 mol of linolenate per mol of albumin respectively indicated that the dye and the fatty acids do not complete for a common primary binding site. In contrast, a secondary Phenol Red-binding site could be identical with the primary octanoate-binding site. Furthermore, the primary Phenol Red-binding site could be the same as a secondary linolenate-binding site. Assignment of the different primary binding sites for Phenol Red and for medium-chain and long-chain fatty acids to a model of the secondary structure of albumin is attempted.  相似文献   

2.
The binding equilibrium of deuteroporphyrin IX to human serum albumin and to bovine serum albumin was studied, by monitoring protein-induced changes in the porphyrin fluorescence and taking into consideration the self-aggregation of the porphyrin. To have control over the latter, the range of porphyrin concentrations was chosen to maker dimers (non-covalent) the dominant aggregate. Each protein was found to have one high-affinity site for deuteroporphyrin IX monomers, the magnitudes of the equilibrium binding constants (25 degrees C, neutral pH, phosphate-buffered saline) being 4.5 (+/- 1.5) X 10(7) M-1 and 1.7 (+/- 0.2) X 10(6) M-1 for human serum albumin and for bovine serum albumin respectively. Deuteroporphyrin IX dimers were found to bind directly to the protein, each protein binding one dimer, with high affinity. Two models are proposed for the protein-binding of porphyrin monomers and dimers in a porphyrin system having both species: a competitive model, where each protein molecule has only one binding site, which can be occupied by either a monomer or a dimer; a non-competitive model, where each protein molecule has two binding sites, one for monomers and one for dimers. On testing the fit of the data to the models, an argument can be made to favour the non-competitive model, the equilibrium binding constants of the dimers, for the non-competitive model (25 degrees C, neutral pH, phosphate-buffered saline), being: 8.0 (+/- 1.8) X 10(8) M-1 and 1.2 (+/- 0.6) X 10(7) M-1 for human serum albumin and bovine serum albumin respectively.  相似文献   

3.
Anion-induced increases in the affinity of colcemid binding to tubulin   总被引:1,自引:0,他引:1  
Colcemid binds tubulin rapidly and reversibly in contrast to colchicine which binds tubulin relatively slowly and essentially irreversibly. At 37 degrees C the association rate constant for colcemid binding is 1.88 X 10(6) M-1 h-1, about 10 times higher than that for colchicine; this is reflected in the activation energies for binding which are 51.4 kJ/mol for colcemid and 84.8 kJ/mol for colchicine. Scatchard analysis indicates two binding sites on tubulin having different affinities for colcemid. The high-affinity site (Ka = 0.7 X 10(5) M-1 at 37 degrees C) is sensitive to temperature and binds both colchicine and colcemid and hence they are mutually competitive inhibitors. The low-affinity site (Kb = 1.2 X 10(4) M-1) is rather insensitive to temperature and binds only colcemid. Like colchicine, 0.6 mol of colcemid are bound/mol of tubulin dimer (at the high-affinity site) and the reaction is entropy driven (163 J K-1 mol-1). Similar to colchicine, colcemid binding to tubulin is stimulated by certain anions (viz. sulfate and tartrate) but by a different mechanism. Colcemid binding affinity at the lower-affinity site of tubulin is increased in the presence of ammonium sulfate. Interestingly, the lower-affinity site on tubulin for colcemid, even when converted to higher affinity in presence of ammonium sulfate, is not recognized by colchicine. We conclude that tubulin possesses two binding sites, one of which specifically recognized the groups present on the B-ring of colchicine molecule and is effected by the ammonium sulfate, whereas the higher-affinity site, which could accommodate both colchicine and colcemid, possibly recognized the A and C ring of colchicine.  相似文献   

4.
Binding equilibria for decanoate to a defatted, commercially available human serum albumin preparation were investigated by dialysis exchange rate determinations. The binding isotherm could not be fitted by the general binding equation. It was necessary to assume that the preparation was a mixture of two albumin components about 40% of the albumin having high affinity and about 60% having low affinity. By affinity chromatography we succeeded in purifying the low-affinity component from the mixture. The high-affinity component, however, could not be isolated. We further analyzed the fatty acid and drug binding abilities of the low-affinity component. The fatty acids decanoate, laurate, myristate and palmitate were bound with higher affinity to the mixture than to the low-affinity component. Diazepam was bound with nearly the same affinity to the low-affinity component as to the albumin mixture, whereas warfarin was not bound at all to the low-affinity component.  相似文献   

5.
The binding of zinc to both human alpha-fetoprotein (AFP) and albumin isolated from cord serum was studied by Sephadex G-50 gel-filtration chromatography. We found that the total number of binding sites for zinc on AFP and albumin were approximately 16 and 12, respectively. Both graphical analysis and the computer program 'LIGAND' indicate that there are at least two major classes of binding sites for both proteins. Both methods of analysis suggested that there are four to five high-affinity sites for zinc on AFP and only two to three similar sites on albumin. The affinity of zinc for AFP (dissociation constant, Kd, 6-8 X 10(-6) mol/l) was higher than for albumin (Kd, 1-3 X 10(-5) mol/l) for the high-affinity sites. The estimates for the zinc low-affinity binding sites were more uncertain, and several classes of low-affinity binding sites of different affinities might be present in both proteins. The results of our inhibition studies suggest that calcium, copper and lead might also bind with AFP at the zinc-binding sites.  相似文献   

6.
1. Several ring-substituted derivatives of diphenyleneiodonium catalyse the exchange of Cl- and OH- ions across the inner membrane of rat liver mitochondria. They also inhibit state 3 and state 3u oxidations of glutamate plus malate in the presence of Cl- more than in its absence. Most have activities similar to diphenyleneiodonium, although 2,4-dichlorodiphenyleneiodonium is up to 50 times more active. 2. Diphenyleneiodonium inhibits soluble rat liver NADH dehydrogenase and NADH oxidation by rat liver sub-mitochondrial particles directly; 2,4-dichlorodiphenyleneiodonium is only about twice as inhibitory. 3. Liver mitochondria contain two classes of binding sites for diphenylene[125I]iodonium, namely high-affinity sites with an affinity constant of 3 X 10(5) M-1 (1--2 nmol/mg of protein), and low-affinity sites with an affinity constant of 1.3 X 10(3) M-1 (80 nmol/mg of protein). Both sites occur in hepatocytes with a relative enrichment of the low-affinity site. Nadh dehydrogenase preparations only apparently contain high-affinity binding sites. Only low-affinity sites occur in erythrocytes. 4. 2,4-Dichlorodiphenyleneiodonium competes with diphenylene[125I]iodonium for both low- and high-affinity sites, whereas tri-n-propyltin only competes for the low-affinity sites. 5. The high-affinity sites are apparently associated with NADH dehydrogenase and the low-affinity sites probably represent electrostatic binding of diphenylene[125I]iodonium to phospholipids. The high-affinity site does not appear to be associated with a rate-limiting stage of NADH oxidation.  相似文献   

7.
The binding of estradiol-17 beta (E2), diethylstilbestrol (DES), and polyene fatty acids, in particular arachidonate (C20:4), to alpha 1-fetoprotein (alpha-FP) and albumin purified from mouse embryo sera was studied using equilibrium dialysis and electrophoretic techniques. E2, arachidonate, and DES all bind to alpha-FP, but with decreasing strength. E2 is a high affinity, low capacity ligand (Ka approximately 0.8 X 10(8) M-1 and approximately 0.3 sites/mol of alpha-FP at 25 degrees C); arachidonate is a weaker ligand disposing of more sites (Ka approximately 0.3 X 10(7) M-1 and 4-5 sites/mol of alpha-FP); the binding of DES is of comparatively low affinity and capacity (Ka approximately 0.2 X 10(7) M-1 and n approximately 0.7/mol of alpha-FP). In spite of different structures and equilibrium parameters, E2, DES, and arachidonate are able to compete with each other for binding to the fetoprotein. The C22:4 and C22:6 fatty acids are also efficient concentration-dependent inhibitors of E2 or DES binding. Albumin binds the fatty acids and DES, but equilibrium parameters are different from those of alpha-FP. In particular, arachidonate is a better ligand for albumin, where it interacts with at least two classes of apparent sites (Ka1 approximately 0.3 X 10(8) M-1 and n1 approximately 1; Ka2 approximately 0.2 X 10(7) M-1 and n2 approximately 30). In contrast to alpha-FP, albumin virtually does not bind E2. Also, no competition could be demonstrated between DES and fatty acid ligands for binding to albumin. None of the studied interactions, with either albumin or alpha-FP, was modified even by high doses of bilirubin. The possible functions of the various binding activities present in fetal sera in the process of growth are discussed.  相似文献   

8.
Excitation-contraction coupling in cardiac muscle is dependent on extracellular calcium and calcium bound to the surface of the myocardial cell. In this study, we examined the physical characteristics of calcium binding to adult guinea pig ventricular myocytes disaggregated mechanically in oxygenated tissue culture medium containing a proteinase inhibitor (aprotinin), and separated from cellular debris by Cytodex beads. Cells prepared in this manner excluded Trypan blue and showed no evidence of spontaneous contraction or contracture. Scatchard plots of calcium binding determined by continuous flow equilibrium dialysis revealed a high-affinity, low-capacity pool, Ka = 65 X 10(3) M-1 and Bt = 1.3 nmol X mg-1 and a low-affinity, high-capacity pool, Ka = 141 M-1 and Bt = 138 nmol X mg-1. The low-affinity pool was not detectable after lanthanum, trypsin or collagenase treatment or in cells prepared without aprotinin in the isolation medium. Both neuraminidase and phospholipase C reduced Bt of the low-affinity pool by one half, but only neuraminidase affected the affinity constant of this pool. Ka was increased to 516.7 M-1, similar to the apparent affinity constant for calcium binding estimated from dP/dtmax measured at several extracellular calcium concentrations (470 M-1). The results suggest that calcium bound to sarcolemmal phospholipids represents the superficial calcium involved in excitation-contraction coupling in the heart.  相似文献   

9.
Vanadate binding to different sarcoplasmic reticulum membrane preparations was determined by measuring bound vanadate colorimetrically and by phosphorylating the vanadate-free enzyme fraction with [gamma-32P] ATP. Colorimetry allowed the study of the dependence of equilibrium vanadate binding on ionized magnesium and the displacing effect of ionized calcium at vanadate concentrations greater than 0.1 mM only. At saturating magnesium concentration the enzyme binds 6-8 nmol vanadate/mg protein and half-maximum saturation is reached at 40 microM. Vanadate is displaced from the enzyme when its high-affinity calcium-binding sites are saturated and conversely calcium is solely displaced from its high-affinity binding sites by vanadate. The phosphorylation procedure allowed the measurement of equilibrium binding as well as the kinetics of vanadate binding and release at vanadate concentrations below 0.1 mM. Half-times of 30s and 3s were observed for vanadate release induced by 0.1 mM and 1 mM calcium respectively. Millimolar concentrations of ATP are required for vanadate displacement. Under equilibrium conditions the enzyme displays an affinity for vanadate of 1.6 X 10(6) M-1. The dependence on the concentration of vanadate of the rate of vanadate binding yielded an affinity of only 1 X 10(4) M-1. Closed vesicles bind vanadate much more slowly than calcium-permeable preparations. The initial rate of calcium-induced vanadate dissociation is accelerated considerably when the vesicles are made calcium permeable. The rate of vanadate dissociation from calcium-permeable vesicles reaches half-maximum values at 1-2 mM calcium indicating that the internal low-affinity calcium-binding sites must first be occupied in order to release bound vanadate. The results suggest that vanadate binding leads to a transition of the external high to internal low-affinity calcium-binding sites.  相似文献   

10.
Calcium titration of the conformational change in cardiac and skeletal troponin C (TN-C) was followed by circular dichroism (CD) at pH values in the range from 5.2 to 7.4. Computer analysis was used to resolve the contributions from the different classes of Ca2+ -binding sites. At pH 6.94 in skeletal TN-C, apparent affinity constants for calcium of 1.8 x 10(7) and 4.5 x 10(5) M-1 were determined for the two classes of binding sites. The more sophisticated computer analysis of the data has revealed a substantial CD contribution from the low-affinity sites (approximately 30% of the high affinity contribution at pH 6.94) and suggests that skeletal TN-C with Ca2+ bound at the low-affinity sites is in a different conformation from that when just the high-affinity sites are occupied, in agreement with a recent nuclear magnetic resonance (NMR) study on this system (Seaman, K. B., Hartshorne, D. J. & Bothener-By, A. A. (1977) Biochemistry 16,4039-4046). With the cardiac protein at pH 7.07, an apparent affinity constant for calcium of 2.0 x 10(7) M-1 was calculated while no low-affinity site at this pH was detected by CD. On the other hand, at lower pH values, such as 6.05, a CD contribution from the cardiac low-affinity Ca2+ -binding site is detected with an apparent binding constant of 3.7 +/- 0.7 x 10(4) M-1. At the lower pH values, protonation of a class of carboxyl groups in each protein which possesses a high pKa (6.2-6.3) elicits the conformational change at the high-affinity sites with a corresponding decrease in the overall magnitude of the Ca2+ -evoked changes. The expression of a conformational change upon Ca2+ binding at the level of the low-affinity sites is enchanced by protonation of a class of carboxyls with a pKa of 6.3 in cardiac TN-C and 6.7-6.8 with the skeletal homologue. In both cases, this contribution is reduced upon protonation of carboxyls with pKa less than or equal to 5.5. It was also observed that the low-affinity sites of skeletal TN-C have a much larger role to play in the total conformational change than the low-affinity sites of cardiac TN-C, a finding probably related to the inability of site 1 in the cardiac protein to bind calcium. In the cardiac protein, the Ca2+ -induced tyrosine difference-spectrum maximum is reduced from deltaepsilonM,287nm =330M-1.cm-1 to 20M-1.cm-1 by protonation of a class of groups with a pKa of 6.4, presumably the same carboxyl groups as those invoved in the CD conformational contribution from the high-affinity binding sites. No such effect was observed for the skeletal protein where deltaepsilonM,287nm was constant at 110M-1 .cm-1 over the pH range studied. The dramatic alterations in the tyrosine environment of cardiac TN-C with pH are attributed to either or both of the tyrosines located in the two high-affinity Ca2+ -binding sites (sites 3 and 4)...  相似文献   

11.
A Miller  rd  S T Li    F Bronner 《The Biochemical journal》1982,208(3):773-781
The Ca2+-binding properties of isolated brush-border membranes at physiological ionic strength and pH were examined by rapid Millipore filtration. A comprehensive analysis of the binding data suggested the presence of two types of Ca2+-binding sites. The high-affinity sites, Ka = (6.3 +/- 3.3) X 10(5) M-1 (mean +/- S.E.M.), bound 0.8 +/- 0.1 nmol of Ca2+/mg of protein and the low-affinity sites, Ka = (2.8 +/- 0.3) X 10(2) M-1, bound 33 +/- 3.5 nmol of Ca2+/mg of protein. The high-affinity site exhibited a selectivity for Ca2+, since high concentrations of competing bivalent cations were required to inhibit Ca2+ binding. The relative effectiveness of the competing cations (1 and 10 mM) for the high-affinity site was Mn2+ approximately equal to Sr2+ greater than Ba2+ greater than Mg2+. Data from the pH studies, treatment of the membranes with carbodi-imide and extraction of phospholipids with aqueous acetone and NH3 provided evidence that the low-affinity sites were primarily phospholipids and the high-affinity sites were either phosphoprotein or protein with associated phospholipid. Two possible roles for the high-affinity binding sites are suggested. Either high-affinity Ca2+ binding is involved with specific enzyme activities or Ca2+ transport across the luminal membrane occurs via a Ca2+ channel which contains a high-affinity Ca2+-specific binding site that may regulate the intracellular Ca2+ concentration and gating of the channel.  相似文献   

12.
Numerical analysis of multiple binding of two ligands to one carrier has been accomplished, using the principle of several sets of acceptable binding constants, with bilirubin-laurate-albumin as an example. Binding of bilirubin to defatted human serum albumin was investigated by a spectroscopic method, based upon a difference of light absorption spectrum for free and bound bilirubin. The observations were supplemented with previous data from an independent technique, measurement of oxidation rates of free bilirubin with hydrogen peroxide and peroxidase. A continuous isotherm was obtained, showing binding of at least 4 mol bilirubin per mole albumin with the following stoichiometric binding constants, 1.11 X 10(8), 1.7 X 10(7), 8 X 10(5), and 4 X 10(4) M-1 at pH 8.2, ionic strength 0.15 M, 25 degrees C. The binding is anticooperative at all steps. A saturation level was not reached. Cobinding of bilirubin and laurate was studied, with up to 2 mol of each ligand per mole albumin, using the peroxidase method for determination of free equilibrium concentrations of bilirubin, and a dialysis rate technique for free laurate. The findings could be described in terms of a stoichiometric model. Heterotropic cooperativity was present among the first bilirubin and the first and second laurate molecules. More than two molecules of either ligand can be bound at the same time.  相似文献   

13.
The interaction of the immobilized triazine dye Cibacron Blue 3G-A with rat, rabbit, sheep, goat, bovine and human serum albumins was studied by affinity gel electrophoresis. Dissociation constants were estimated in each instance and showed human serum albumin to have a significantly higher affinity for the dye than did albumin from any other species. Pretreatment of the defatted proteins with bilirubin (3 mol of bilirubin/mol of protein) did not increase the dissociation constants of the serum albumins, whereas pretreatment with palmitate (7 mol of palmitate/mol of protein) increased the dissociation constant in all cases: 3-fold for human serum albumin, 15-fold for other serum albumins. Increasing the bilirubin/albumin ratio (to 7:1) did not affect the dissociation constant of the albumins studied. Decreasing the palmitate/albumin ratio decreased the dissociation constant for human serum albumin, but did not affect those of bovine and rat albumins. Altering the chain length of the presaturating fatty acid dramatically changed the dissociation constant of both human and bovine serum albumins. Butyrate, hexanoate, octanoate and decanoate did not significantly influence the dissociation constants of bovine and human serum albumins for Cibacron Blue, whereas laurate, myristate and palmitate greatly increased the dissociation constant. These data are discussed in relationship to the behaviour of albumins during dye--agarose column chromatography. In Addendum the effect of nucleotide presaturation on the interaction between Bacillus stearothermophilus 6-phosphogluconate dehydrogenase and the immobilized triazine dyes Cibacron Blue 3G-A and Procion Red HE-3B was examined, and the implications for dye--ligand chromatography are discussed.  相似文献   

14.
A thyroid hormone binding protein(s) has been characterized in the cytosol of fetal rat brain cells in primary cultures. This protein is closely related to the one described in brain supernatants with respect to its electrophoretic mobility, binding kinetic parameters and estimated molecular weight (65 000 daltons). However, in contrast to the brain cytosolic binding protein, two classes of affinity sites for triiodothyronine (T3) and thyroxine (T4) have been demonstrated: a high affinity site (KA = 1.2-3.7(3) X 10(9) M-1 for T3 and KA = 3.7-5 X 10(8) M-1 for T4) and a low affinity site (KA = 0.8-1.4 X 10(8) M-1 for T3 and 1.6-2.9 X 10(7) M-1 for T4). The results are discussed with respect to their cellular significance.  相似文献   

15.
Plasma protein binding interaction of prednisone and prednisolone   总被引:2,自引:0,他引:2  
The plasma protein binding interaction of prednisone and prednisolone were characterized by equilibrium dialysis. Prednisone and prednisolone are bound equally but weakly to human albumin (affinity constant, K approximately equal to 1 X 10(3) M-1). Transcortin affinity for prednisolone is 10-fold greater (51.3 X 10(6) M-1) than that for prednisone (4.3 X 10(6) M-1). In competition under pharmacologic conditions, prednisolone inhibits prednisone binding to transcortin producing linear binding averaging 55%. Prednisone does not affect prednisolone binding and does not complicate pharmacokinetic studies of the latter.  相似文献   

16.
The calcium binding properties of non-activated phosphorylase kinase at pH 6.8 have been studied by the gel filtration technique at calcium concentrations from 50 nM to 50 muM. Taking into account the subunit structure alpha4beta4gamma4 the enzyme binds 12 mol Ca2+ per mol with an association constant of 6.0 X 10(7) M-1, 4 mol with an association constant of 1.7 X 10(6) M-1 and 36 mol with a binding constant of 3.9 X 10(4) M-1 at low ionic strength. In buffer of high ionic strength, i.e. 180 mM NH4Cl or 60 mM (NH4)2SO4, only a single set of eight binding sites with a binding constant of 5.5 X 10(7) M-1 is left. In a buffer containing 155 mM NH4Cl and 10 mM MgCl2, the calcium affinity of these sites is reduced to a KCa of 3.0 X 10(6) M-1, indicating competition between Ca2+ and Mg2+. From these measurements, the binding constant of Mg2+ for these sites is calculated to be 1.7 X 10(3) M-1 is left. In a buffer containing 155 mM NH4Cl and 10 mM MgCl2, the calcium affinity of these sites is reduced to a KCa of 3.0 X 10(6) M-1, indicating competition between Ca2+ and Mg2+. from these measurements, the binding constant of Mg2+ for these sites is calculated to be 1.7 X 10(3) M-1. Additionally, 10 mM Mg2+ induces a set of four new Ca2+ binding sites which show positive cooperativity. Their half-saturation constant under the conditions described is 3.5 X 10(5) M-1, and they, too, exhibit competition between Ca2+ and Mg2+. Since this set of sites is induced by Mg2+ a third group of binding sites for the latter metal must be postulated.  相似文献   

17.
Insulin receptors of rat skeletal muscle were purified by first extracting a plasma membrane-enriched pellet obtained from a muscle homogenate with Triton X-100, followed by WGA-Sepharose and insulin-Sepharose affinity chromatography. Routinely, 4-5 micrograms of purified receptor were obtained from 15 g of tissue. The purified receptors are composed of two major polypeptides with molecular weights of 130,000 and 95,000, respectively. The binding of [125I]insulin by the purified receptors was analyzed by a Scatchard plot. There are at least two binding components. The high-affinity component, with an apparent association constant (Ka) of 2.0 X 10(9) M-1, comprises 10% of the total insulin binding sites; while the low-affinity component, with a Ka value of 1.4 X 10(8) M-1, represents 90% of the binding sites. Assuming the insulin receptor to have a molecular weight of 300,000, the receptor binds 1.7 mol of insulin per mol at saturation. Insulin is capable of stimulating the autophosphorylation of the beta-subunit of the muscle insulin receptor (Mr 95,000) by 5-10-fold. The stoichiometry of this phosphorylation reaction was determined as 0.8 phosphate per insulin binding site after a 10 min incubation with 100 nM insulin. In a previous report, I showed that the insulin stimulation of glucose transport in diaphragms from neonatal rats was small, even although the diaphragms had normal levels of insulin receptors and glucose transporters (Wang, C. (1985). Proc. Natl. Acad. Sci. USA 82, 3621-3625). To determine whether or not receptor autophosphorylation might be related to this insensitivity to insulin, the level of receptor phosphorylation was quantitated in diaphragms from rats at different stages of development. Autophosphorylation remains unchanged from birth to 21 days of age, suggesting that the lower insulin-stimulated glucose uptake by diaphragms at early stages of postnatal development as compared to that by diaphragms of older rats, is not due to a difference in receptor kinase.  相似文献   

18.
The binding of [14C]ellipticine to native calf thymus DNA was studied using equilibrium dialysis. A Scatchard polt revealed the presence of high-and low-affinity binding sites in DNA, the former having a K of 4.0 X 10(7) M(-1) and an n (saturation limiting of binding) of 0.078 (1mol ellipticine/13 mol of DNA nucleotides). The forces involved in stabilizing the high-affinity binding, which has been equated with intercalative binding, were due to a combination of hydrophobic interactions and hydrogen bonding. Difference spectra of ellipticine in the presence of the polydeoxynucleotides, poly d(A-T) or poly d(G-C), showed that there was no base specificity involved in the high-affinity binding. Ellipticine binding to the low-affinity sites, which has been equated with surface binding, was due primarily to the participation of electrostatic interactions of ellipticine with the anionic phosphate groups on the double helical surface of DNA.  相似文献   

19.
Bovine serum albumin (BSA) is routinely utilized in vitro to prevent the adverse detergent effects of long-chain acyl-CoA esters (i.e., palmitoyl-CoA) in enzyme assays. Determination of substrate saturation kinetics in the presence of albumin would only be valid if the relationship between bound and free substrate concentrations was known. To elucidate the relationship between bound and free palmitoyl-CoA concentrations in the presence of BSA, several different techniques including equilibrium dialysis, equilibrium partitioning, fluorescence polarization and direct fluorescence enhancement were investigated. Direct fluorescence enhancement using a custom synthesized fluorescent probe, 16-(9-anthroyloxy)palmitoyl-CoA (AP-CoA), was the best approach to this question. Measurement of the relationship between mol of palmitoyl-CoA bound per mol of BSA (nu) versus -log[free palmitoyl-CoA] revealed that the binding of palmitoyl-CoA to BSA, like palmitate was nonlinear, suggesting the presence of more than one class of acyl-CoA binding sites. Computer analyses of the binding data gave a best fit to the 2,4 two-class Scatchard model, suggesting the presence of two high-affinity primary binding sites (k1 = (1.55 +/- 0.46) x 10(-6) M-1) and four lower affinity secondary binding sites (k2 = (1.90 +/- 0.09) x 10(-8) M-1). Further analyses using the six parameter stoichiometric (stepwise) ligand binding model supports the existence of six binding sites with the higher affinities associated with the binding of the first mole of palmitoyl-CoA and weaker binding occurring after the first two sites are occupied. The association constants from this model of multiple binding diminish sequentially (i.e., K1 greater than K2 greater than K3 greater than...greater than or equal to K6), suggesting that each mol of long-chain acyl-CoA binds to BSA with decreasing affinities.  相似文献   

20.
The specific binding of IgG to jejunal brush borders was greatest at acidic pH, at neutral pH no specific binding occurred. Specific binding declined with age-no specific binding occurred in borders from 20-and 24-day-old animals. There was no specific binding of IgG to borders from ileal enterocytes. Human transferrin and bovine serum albumin did not bind specifically to borders. The affinity of binding (-Ka) and the receptors site numbers per border estimated for rat IgG were 18.64 X 10(6) M-1 to 3.53 X 10(6) sites; for human IgG, 25.06 X 10(6) M-1 to 3.30 X 10(6) sites; for bovine IgG, 10.48 X 10(6) M-1 to 2.11 X 10(6) sites and for sheep IgG, 7.26 X 10(6) M-1 to 2.34 X 10(6) sites.  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