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.
Centrifuge transport, equilibrium dialysis, and electron paramagnetic resonance studies on the binding of Mn2+ to myosin revealed two sets of noninteracting binding sites which are characterized at low ionic strength (0.016 M KCl) by affinity constants of 10(6) M-1 (Class I) and 10(3) M-1 (Class II), respectively. At 0.6 M KCl concentration, the affinity of Mn2+ for both sets of sites is reduced. The maximum number of binding sites is 2 for the high affinity and 20 to 25 for the low affinity set. Other divalent metal ions displace Mn2+ from the high affinity sites in the following order of effectiveness: Ca greater than Mg = Zn = Co greater than Sr greater than Ni. The inhibitory effects of Mg2+ and Ca2+ upon the Mn2+ binding are competitive with inhibitor constants of 0.75 to 1 mM which is similar to that of the low affinity divalent metal ion binding sites. Exposure of myosin to 37 degrees partially inhibits Mn2+ binding to Class I parallel with inhibition of ATPase activity. The binding of Mn2+ to the high affinity binding sites is not significantly influenced by ADP or PPi, although Mn2+ increases the affinity of ADP binding to myosin at high ionic strength.  相似文献   

2.
Y Doi  F Kim  S Kido 《Biochemistry》1990,29(6):1392-1397
Calcium binding of swine plasma gelsolin was examined. When applied to ion-exchange chromatography, its elution volume was drastically altered depending on the free Ca2+ concentration of the medium. The presence of two classes of Ca2+ binding sites, high-affinity sites (Kd = 7 microM) and low-affinity sites (Kd = 1 mM), was suggested from the concentration dependence of the elution volume. The tight binding sites were specific for Ca2+. The weakly bound Ca2+ could be replaced by Mg2+ once the tight binding sites were occupied with Ca2+. The binding of metal ions was totally reversible. Circular dichroism measurement of plasma gelsolin indicated that most change in secondary structure was associated with Ca2+ binding to the high-affinity sites. Binding of Mg2+ to the low-affinity sites caused a secondary structural change different from that caused by Ca2+ bound to the high-affinity sites. Gel permeation chromatography exhibited a small change in Stokes radius with and without Ca2+. Microheterogeneity revealed by isoelectric focusing did not relate to the presence of two classes of Ca2+ binding sites. These results indicated that plasma gelsolin drastically altered its surface charge property due to binding of Ca2+ or Ca2+, Mg2+ with a concomitant conformational change.  相似文献   

3.
The broadening of spin-label absorption lines resulting from spin-exchange reactions that occur during collision with paramagnetic Ni2+ is diminished when Ni2+ binds to phospholipid vesicles. Subsequent addition of non-paramagnetic ions that compete for binding sites releases Ni2+ into solution and restores the line-broadening. The concentrations of various ions required to achieve this effect was used to order the ions with respect to their binding to vesicles containing phosphatidylethanolamine and phosphatidylglycerol. The relative strengths of binding for those ions studied were: Ca2+ > Mg2+ > Zn2+ > Sr2+ > Ba2+. The spin-broadening assay was also used to study the effects of two proteins on the availability of Ni2+-binding sites on the vesicles. Ribonuclease, which is thought to associate electrostatically as an extrinsic protein on the surface of vesicles, completely blocked the Ni2+-binding sites at comparatively low protein concentrations. Quantitative considerations of these data suggest the possibility that Ni2+ may bind preferenetially to phosphatidylglycerol, and that these binding sites are aggregated in the ribonuclease-containing vesicles. In contract to ribonuclease, cytochrome c does not block Ni2+-bindings sites on the phospholipid vesicles, but rather contains sites of its own that bind Ni2+, both when the protein is in solution and when it is associated with the vesicles. These results are consistent with other studies which suggest that cytochrome c becomes partially embedded in membrane bilayers and associates with phospholipid molecules through hydrophobic interactions.  相似文献   

4.
Conformational changes induced by binding of divalent cations to calregulin   总被引:3,自引:0,他引:3  
Scatchard analysis of equilibrium dialysis studies have revealed that in the presence of 3.0 mM MgCl2 and 150 mM KCl, calregulin has a single binding site for Ca2+ with an apparent dissociation constant (apparent Kd) of 0.05 microM and 14 binding sites for Zn2+ with apparent Kd(Zn2+) of 310 microM. Ca2+ binding to calregulin induces a 5% increase in the intensity of intrinsic fluorescence and a 2-3-nm blue shift in emission maximum. Zn2+ binding to calregulin causes a dose-dependent increase of about 250% in its intrinsic fluorescence intensity and a red shift in the emission maximum of about 11 nm. Half-maximal wavelength shift occurs at 0.4 mol of Zn2+/mol of calregulin, and 100% of the wavelength shift is complete at 2 mol of Zn2+/mol of calregulin. In the presence of Zn2+ and calregulin the fluorescence intensity of the hydrophobic fluorescent probe 8-anilino-1-napthalenesulfonate (ANS) was enhanced 300-400% with a shift in emission maximum from 500 to 480 nm. Half-maximal Zn2+-induced shift in ANS emission maximum occurred at 1.2 mol of Zn2+/mol of calregulin, and 100% of this shift occurred at 6 mol of Zn2+/mol of calregulin. Of 12 cations tested, only Zn2+ and Ca2+ produced changes in calregulin intrinsic fluorescence, and none of these metal ions could inhibit the Zn2+-induced red shift in intrinsic fluorescence emission maximum. Furthermore, none of these cations could inhibit or mimic the Zn2+-induced blue shift in ANS emission maximum. These results suggest that calregulin contains distinct and specific ligand-binding sites for Ca2+ and Zn2+. While Ca2+ binding results in the movement of tryptophan away from the solvent, Zn2+ causes a movement of tryptophan into the solvent and the exposure of a domain with considerable hydrophobic character.  相似文献   

5.
Small unilamellar vesicles have been prepared from phosphatidylethanolamine by sonication of the lipid in aqueous buffers of low ionic strength and high pH. These vesicles and their interactions with various di- and trivalent cations have been characterized using freeze-fracture electron microscopy. Phosphatidylethanolamine from 4 sources was examined: Hens' yolk phosphatidylethanolamine, human grey matter phosphatidylethanolamine, Escherichia coli phosphatidylethanolamine and dimyristoyl phosphatidylethanolamine. The phosphatidylethanolamine from natural sources formed spherical, uniform 20–40 nm vesicles while dimyristoyl phosphatidylethanolamine formed larger, 70 × 25 nm, disc-shaped vesicles when sonicated above the phase transition temperature. Fusion of the unilamellar egg phosphatidylethanolamine, E. coli phosphatidylethanolamine and human grey matter phosphatidylethanolamine vesicles was induced by dialysis against buffers containing 2.0 nM Ca+ or 3.0 mM Mg2+. The fusion of the vesicles resulted in the precipitation of the lipid and the formation of multilamellar and, in some cases, hexagonal II structures. Dimyristoyl phosphatidylethanolamine vesicles were precipitated at 55°C by 1.0 mM Ca+ or 2.0 mM Mg2+. Treatment of the calcium- and magnesium-precipitated vesicles of hen's egg yolk phosphatidylethanolamine, E. coli phosphatidylethanolamine, human grey matter phosphatidylethanolamine and dimyristoyl phosphatidylethanolamine with EDTA resulted in resuspension of the lipid. The specific size and shape of the vesicles formed in this manner depends on the type of phosphatidylethanolamine and ion involved. Dialysis of the Ca+- and Mg2+-precipitated egg phosphatidylethanolamine vesicles against buffer containing no Ca+, Mg2+ or EDTA also resulted in dissociation of the precipitate and formation again of a new vesicle population. This evidence indicates that the Ca+ and Mg2+ are not strongly bound to the phosphatidylethanolamine.Egg phosphatidylethanolamine vesicles would fuse in the presence of many di- and trivalent ions. Egg phosphatidylethanolamine vesicles were precipitated by beryllium, aluminum, chromium, manganese, cobalt, nickel, copper, zinc, strontium, cadmium, barium, lanthanium, mercury and lead. The amount of ion required to precipitate the vesicles and the type of structure resulting from the fusion of the vesicles was found to be unique for each ion.Small unilamellar vesicles prepared from egg phosphatidylethanolamine were reacted with several basic proteins (cytochrome c, basic protein from human myelin, protamine, poly-l-lysine and cationically-modified ferritin). The basic proteins also initiated the fusion of egg phosphatidylethanolamine vesicles but these proteins did not fuse egg phosphatidylcholine vesicles nor did normal ferritin initiate fusion. Human myelin basic protein initiated the fusion of dimyristoyl phosphatidylethanolamine vesicles above and below the phase transition of this lipid.  相似文献   

6.
We have used the osmotic pressure technique of Rand, Parsegian and co-workers (Nature 259 (1976) 601–603) to investigate the effect of anion species on the binding of M2+ to dipalmitoylphosphatidylcholine bilayers. Calcium and magnesium salts show a complex behavior which is consistent with both anion binding and screening. We observe virtually no change, within the accuracy of our experiment, in the decay of repulsive pressure with inter-bilayer separation for the acetate and nitrate salts of magnesium and calcium; however, the chloride salt does show a different pressure decay. At any given bilayer separation, , with calcium and magnesium salts present, the anions produce a decrease in the repulsive pressure in the order acetate > Cl > NO3.  相似文献   

7.
Dynamics of ligand binding to heme proteins   总被引:23,自引:0,他引:23  
  相似文献   

8.
P Argos  M G Rossmann 《Biochemistry》1979,18(22):4951-4960
Of the 82 three dimensionally characterized residues of cytochrome c551, 49 are found to be structurally and topologically equivalent to the globin fold and 41 are equivalent to the cytochrome b5 fold, with a respective root mean square separation of 3.5 and 4.9 A between equivalenced Calpha atoms. The common fold represents a central heme binding core, corresponding to the middle exon of certain globin genes. After superposition of the protein folds, the heme irons are found to be separated by 5.4 and 1.6 A, while their heme normals are inclined by 6 degrees and 32 degrees, respectively. Furthermore, the heme "face", determined by the asymmetric attachment of the vinyl and propionyl side chains, is directed similarly in all three heme proteins. The heme itself is rotated by 72 degrees and 116 degrees about its normal, respectively. The minimum base change per codon for the three pairwise comparisons corresponds to the expected value of random sequence comparisons. While all three heme proteins may have diverged from a common ancestor, their similarity may have arisen from the requirements of heme binding or the utilization of a particularly stable fold. Known structures within commonly accepted divergent families were superimposed in order to discriminate better between convergence and divergence. Minimum base changes per codon, number of deletions and insertions, percentage of equivalenced residues, precision of heme superposition, and root mean square separation of equivalenced Calpha atoms were tested as measures of evolutionary relationships.  相似文献   

9.
10.
Solubilized membrane proteins from HL-60 cells were separated by two-step affinity chromatography. Proteins eluted with MgCl2 in the first heparin-gel were applied to the second heparin-gel and eluted with CaCl2. The eluted proteins were analysed and purified by electrophoresis. N-terminal amino acid sequences of eight proteins on the characteristic bands were determined. Homology search for the sequences indicated that three microsomal proteins, two nuclear proteins and a glycolytic enzyme were eluted with divalent cations, whereas a nuclear ribonucleoprotein and a membrane-cytoskelton linker protein were not dissociated with divalent cations, but with 2 M NaCl. Heparin affinity chromatography combined with differential elution with divalent cations can be a useful method for separation of membrane proteins.  相似文献   

11.
12.
K Grizzuti  G E Perlmann 《Biochemistry》1975,14(10):2171-2175
Dialysis equilibrium measurements at 25 degrees indicate that, at pH 6.8 and at a concentration of 1.0 times 10(-10) 3 M MnC12 or CoC12, phosvitin binds 113 Mn2+ and 120 Co2+. The binding is cooperative at low cation concentrations. The number of Mg2+, Ca2+, Mn2+, and Co2+ bound is not affected by temperatures of up to 60 degrees; however, the cooperactivity is enhanced. Optical rotatory dispersion and circular dichroism studies indicate that a conformational change occurs on binding of Mn2+ and Co2+ which parallels the one produced by Ca2+ and reported elsewhere [Grizzuti, K., and Perlmann, G.E. (1973), Biochemistry 12, 4399]. The conformational changes induced by Mg2+ and Mn2+ follow different paths. Upon binding of Mn2+ and Co2+ the intrinsic viscosity, [eta], of phosvitin decreases from about 0.5 to 0.03 dl/g, while Mg2+ and Ca2+ decrease [eta] to 0.048 dl/g. The ultraviolet absorption spectrum of phosvitin is altered upon binding of Ca2+, Mn2+, and Co2+, but not upon binding of Mg2+; an increase of the temperature to 60% has no further effect on the spectra.  相似文献   

13.
Of the 85 three-dimensionally characterized residues of cytochrome b5, 51 are structurally and topologically equivalent to the globin fold. When these proteins have been superimposed, the heme irons are found to be less than 1.4 A separated and the heme normals are inclined by less than 9.5 degrees. Comparison of minimum base changes per codon between heme binding and NAD binding proteins are of the same order.  相似文献   

14.
The effects of zinc and other divalent metals on the [125I]T3 binding to rat c-erbA alpha and beta recombinant proteins were assessed. The addition of ZnCl2 caused a reversible and dose-dependent inhibition of [125I]T3 binding to rc-erbA beta proteins with half maximum inhibition occurring at 50-100 microM, but no significant effect on [125I]T3 binding to rc-erbA alpha under the same assay conditions. Scatchard analysis revealed a decrease in [125I]T3 binding capacity to beta protein without marked change in Kd values in presence of zinc. Moreover, significant inhibitions of [125I]T3 binding to both alpha and beta proteins were observed in the presence of 100 microM of either MnCl2, CdCl2 or CuCl2, but not MgCl2. Thus, the selective effect of zinc compared to other divalent metals to inhibit T3 binding to rc-erbA beta, but not alpha, proteins was documented and suggest a possible regulatory role for zinc in modulating the intracellular action of thyroid hormone.  相似文献   

15.
Binding of chromomycin A3 (CRA) to calf thymus DNA was investigated in the presence of divalent cations using visible absorption and 1H-nmr spectroscopies. An apparent equilibrium binding constant (approximately 10(11) M-1) was obtained from metal competition experiments using EDTA to remove the metal cation from the DNA-M-CRA (M: metal) complex. The large binding constant of the drug to DNA enabled us to obtain essentially complete complexation of CRA to the short homogeneous d(ATGCAT)2 duplex using stoichiometric amounts of the metal cation. Large induced chemical shifts were observed in the 1H-nmr spectrum of the above complex using the paramagnetic Co2+ cation, indicating that the metal occupies a unique binding site. Since no induced 1H-nmr chemical shifts were observed for the drug-Co2+ mixture, it was concluded that no metal-drug complex is formed. In addition, it was found that bound CRA is negatively charged at physiological pH and binding to the DNA could be affected only by using metal cations whose ionic radius size (less than 0.85 A) and charge (2+) were simultaneously satisfied. Stringent metal cation selectivity for the DNA-M-CRA complex may be intimately connected with the antitumor selectivity of CRA, since different types of cells generally possess widely differing molar concentrations of metal cations.  相似文献   

16.
Microfibrils and elastin are major constituents of elastic fibers, the assembly of which is dictated by multimolecular interactions. Microfibril-associated glycoprotein-1 (MAGP-1) is a microfibrillar component that interacts with the soluble elastin precursor, tropoelastin. We describe here the adaptation of a solid-phase binding assay that defines the effect of divalent cations on the interactions between MAGP-1 and tropoelastin. Using this assay, a strong calcium-dependent interaction was demonstrated, with a dissociation constant of 2.8 +/- 0.3 nm, which fits a single-site binding model. Manganese and magnesium bestowed a weaker association, and copper did not facilitate the protein interactions. Three constructs spanning tropoelastin were used to quantify their relative contributions to calcium-dependent MAGP-1 binding. Binding to a construct spanning a region from the N-terminus to domain 18 followed a single-site binding model with a dissociation constant of 12.0 +/- 2.2 nm, which contrasted with the complex binding behavior observed for fragments spanning domains 17-27 and domain 27 to the C-terminus. To further elucidate binding sites around the kallikrein cleavage site of domains 25/26, MAGP-1 was presented with constructs containing C-terminal deletions within the region. Construct M1659, which spans a region from the N-terminus of tropoelastin to domain 26, inclusive, bound MAGP-1 with a dissociation constant of 9.7 +/- 2.0 nm, which decreased to 4.9 +/- 1.0 nm following the removal of domain 26 (M155n), thus displaying only half the total capacity to bind MAGP-1. These results demonstrate that MAGP-1 is capable of cumulative binding to distinct regions on tropoelastin, with different apparent dissociation constants and different amounts of bound protein.  相似文献   

17.
It has been repeatedly postulated that the high heat resistance of bacterial spores is due to stabilization of biopolymers in the spore interior by a solid deposit of protective cement consisting of coordination complexes of ligands with divalent metal ions. This report presents data on metal-binding characteristics of some of the ligands related to spores as determined by means of potentiometric equilibrium measurements under conditions of temperature and ionic strength (t = 25.0°C; μ = 1.0 KNO3) identical with those reported earlier by the authors in order to facilitate correlation by using comparable data. The spore ligands investigated in this study included 2,6-pyridinedicarboxylic acid (DPA), α,ε-diaminopimelic acid, D-glutamic acid, and D-alanine in a ratio of 1:1 with metal ions which are known to play a role in heat resistance of spores. Stability constants of the chelates of these spore ligands with metal ions such as Ca(II), Mg(II), Cu(II), Ni(II), Zn(II), Co(II), and Mn(II) have been determined. In general the metal chelates of DPA exhibited the greatest stability. On the basis of a consideration of the stability data together with the known configurations of the ligand and the coordination requirements of the metal ions, possible structures indicating the coordinate binding of the spore ligands with the metal ions are presented. All the metal chelates except those of Ca(II) were found to undergo hydrolysis and separation of solid phase in the pH range 7-8.5. The relatively greater hydrolytic stability of Ca(II) chelates and the high affinity of DPA for metal ions appear to be of biological significance insofar as these two spore components are more widely associated with the heat resistance of bacterial spores.  相似文献   

18.
The distribution of divalent ions in semidilute solutions of high-molecular-mass DNA containing both sodium chloride and strontium chloride in near-physiological conditions is studied by small-angle x-ray scattering and by small-angle neutron scattering. Both small-angle neutron scattering and small-angle x-ray scattering reveal a continuous increase in the scattering intensity at low q with increasing divalent ion concentration, while at high q the scattering curves converge. The best fit to the data is found for a configuration in which DNA strands of cross-sectional radius 10 angstroms are surrounded by a counterion sheath of outer radius approximately 13.8 angstroms, independent of the strontium chloride concentration. When the strontium chloride is replaced by calcium chloride, similar results are obtained, but the thickness of the sheath increases when the divalent salt concentration decreases. These results correspond in both cases to partial localization of the counterions within a layer that is thinner than the effective Debye screening length.  相似文献   

19.
The binding of citrate and acidic peptide DDSDEEN with DNA in the presence of divalent cations is compared. Citric acid shows a higher number of binding sites on the DNA compared to the peptide; this is probably due to the bigger sitric hindrance of the peptide compared to the citric acid for the binding in the DNA grooves. Moreover, DNA preincubated with saturating amounts of citric acid is not available for the binding with successively added peptide. Therefore the peptide and citrate binding sites to some extent overlap.  相似文献   

20.
Selective binding of cholesterol by recombinant fatty acid binding proteins   总被引:3,自引:0,他引:3  
The sterol binding specificity of rat recombinant liver fatty acid binding protein (L-FABP) and intestinal fatty acid binding protein (I-FABP) was characterized with [3H]cholesterol and a fluorescent sterol analog dehydroergosterol. Ligand binding analysis, fluorescence spectroscopy, and activation of microsomal acyl-CoA:cholesterol acyltransferase activity showed that L-FABP-bound sterols. 1) Lipidex-1000 assay showed a dissociation constant Kd = 0.78 +/- 0.18 microM and stoichiometry of 0.47 +/- 0.16 mol/mol for [3H]cholesterol binding to L-PABP. 2) With [3H]cholesterol/phosphatidylcholine liposomes, the cholesterol binding parameters for L-FABP were Kd = 1.53 +/- 0.28 microM and stoichiometry 0.83 +/- 0.07 mol/mol. 3) L-FABP interaction with dehydroergosterol altered the fluorescence intensity and polarization of dehydroergosterol. Dehydroergosterol bound to L-FABP with Kd = 0.37 microM and a stoichiometry of 0.83 mol/mol. 4) Cholesterol and dehydroergosterol decreased L-FABP tyrosine lifetime. Dehydroergosterol binding produced sensitized emission of bound dehydroergosterol with longer lifetime.5) L-FABP bound two cis-parinaric acid molecules/molecule of protein. Cholesterol displaced one of these bound cis-parinaric acids. 6) L-FABP enhanced acyl-CoA:cholesterol acyltransferase in a concentration-dependent manner. In contrast, these assays indicated that I-FABP did not bind sterols. Thus, L-FABP appears able to bind 1 mol of cholesterol/mol of L-FABP, the L-FABP sterol binding site is equivalent to one of the two fatty acid binding sites, and L-FABP stimulates acyl-CoA:cholesterol acyltransferase by transfer of cholesterol.  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