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.
Al(III) toxicity in living organisms is based on competition with other metal cations. Mg(II) is one of the most affected cations, since the size similarity dominates over the charge identity. The slow ligand exchange rates for Al(III) render the ion useless as a metal ion at the active sites of enzymes and provide a mechanism by which Al(III) inhibits Mg(II) dependent biochemical processes. Al(III) cation interactions with relevant bioligands have been studied in a protein-model environment in gas and aqueous phases using density functional theory methods. The protein model consists of the metal cation bound to two chosen bioligands (functional groups of the amino acid side chains, one of them being always an acetate) and water molecules interacting with the cation to complete its first coordination shell. Analogous Mg(II) complexes are calculated and compared with the Al(III) ones. Formation energies of the complexes are calculated in both phases and magnesium/aluminum exchange energies evaluated. The effect of different dielectric media is also analyzed. The presence of an acetate ligand in the binding site is found to promote both, complex formation and metal exchange reactions. In addition, buried binding sites (with low dielectric constant) of the protein favor metal exchange, whereas fully solvated environments of high dielectric constant require the presence of two anionic ligands for metal exchange to occur.  相似文献   

2.
The aluminum (III) complex [SalenAl(III)]Cl (1), (Salen=(R,R)-N,N'-bis[5-methyl-3-(4-methylpiperazinyl)-salicylidene]-1,2-diphenylethanediamine) has been synthesized and characterized by elemental analysis, FT-IR, (1)H and (13)C NMR measurements. The interaction of complex (1) with calf thymus (CT) DNA has been studied extensively by experimental techniques. Thermal denaturation study of DNA with (1) revealed the DeltaT(m) of 5+/-0.2 degrees C. Viscosity and steady-state fluorescence measurements showed that the complex cross-links DNA and the metal center is interacting with DNA during the cross-linking. Also, the phenyl ring in the complex may intercalate between the base pairs of the DNA during the cross-linking. Competitive binding study shows that the enhanced emission intensity of ethidium bromide (EB) in the presence of DNA was quenched by the addition of the metal complex indicating that it displaces EB from its binding site in DNA and the apparent binding constant has been estimated to be (2.8+/-0.2)x10(5) M(-1). Further, time-resolved fluorescence experiments confirm the binding of (1) with DNA and its cross-linking nature. Aluminum ions shown to precipitate DNA completely above the pH 6.0, but no such precipitation was observed with complex (1). The DNA-DNA cross-linking mediated by (1) is further confirmed by gel electrophoresis.  相似文献   

3.
Competition of the trivalent cation, Co(NH3)(3+)(6), with K+ and Na+ ions in binding to DNA was studied by equilibrating oriented DNA fibers with ethanol/water solutions (65 and 52% v/v EtOH), containing different combinations and concentrations of KCl and NaCl and constant concentration (0.8 mM) of Co(NH3)(6)Cl(3). The degree of Co(NH3)(3+)(6) binding to DNA does not depend significantly on the ethanol concentration or on the kind of univalent cation (Na+ or K+). The ion exchange selectivity coefficient of monovalent-trivalent ion competition, D(1)(c3), increases with the concentration of Me+, C(o)(+), and the monotonic dependence of log D(1)(c3) vs log C(o)(+) has an inflection between 100 and 300 mM that is caused by a structural transformation of DNA from A- to B-form. The ion exchange experimental data are compared with results of grand canonical Monte Carlo (GCMC) simulations of systems of parallel and hexagonally ordered, discretely charged polyions with density and spatial distribution of the charged groups modeling B- and A-forms of DNA. The GCMC method for discretely charged models of the DNA polyion produces a quantitative agreement with experimental data on trivalent-monovalent ion competition in dependence on DNA structural state and salt concentration. Based on this and previous studies it is concluded that the affinity of DNA for the cations decreases in the order Co(NH3)(3+)(6) > Ca2+ > Mg2+ > Na+ approximately K+ > Li+. DNA does not exhibit selectivity for Na+ or K+ in ethanol/water solutions either in the absence or in the presence of Co(NH3)(3+)(6), Ca2+, and Mg2+.  相似文献   

4.
The electron transfer reaction between ferrocyanide ion and the blue copper protein, stellacyanin, has been investigated by means of 13C NMR line broadening of the inorganic oxidant. The temperature dependence of the ferrocyanide line broadening gives an activation energy for the electron transfer reaction of 17 +/- 3 kJ. The apparent rate constant decreases with increasing concentration of K4Fe(CN)6, a result which can be explained either by formation of a strong precursor ferrocyanide--stellacyanin [Cu(II)] complex or by increased formation of KFe(CN)3-6 ion pairs. The direct electron transfer between ferrocyanide and ferricyanide has also been studied by 13C NMR line broadening of the former species. The ferricyanide concentration dependence of the exchange line broadening yields a value for the apparent second-order rate constant at 25 degrees C of k = 1.65 . 10(3) M-1 . s-1, in agreement with previously reported values derived from 14N NMR and isotope exchange studies. This rate constant shows a linear dependence on the K+ concentration, independent of ionic strength, a result which confirms the importance of ion pair species such as KFe(CN)3-6 and KFe(CN)2-6 in the direct electron transfer mechanism. The general applications of the method are discussed, including the considerations which suggest that a wide range of electron transfer rates, from about 1 s-1 to 4 . 10(3) s-1, are, in principle, accessible to this technique. The potential utility of ferrocyanide 13C spin--lattice relaxation time measurements is decreasing the lower limit of this range is also discussed.  相似文献   

5.
Cr(III) and Cr(VI) have different binding capacity with sparfloxacin, and have different combination modes with calf thymus DNA. Selecting these two different metal ions, the influence of them on the binding constants between sparfloxacin (SPFX) and calf thymus DNA, as well as the related mechanism has been studied by using absorption and fluorescence spectroscopy. The result shows that Cr(III) has weaker binding capacity to SPFX in the SPFX-Cr(III) binary system, but influences the binding between SPFX and DNA obviously in SPFX-DNA-Cr(III) ternary system. However, although Cr(VI) has a stronger binding capacity to SPFX, it has no effect on the binding between SPFX and DNA. Referring to the different modes of Cr(III) and Cr(VI) binding to DNA, the mechanism of the influence of metal ions on the binding between SPFX and DNA has been proposed. SPFX can directly bind to DNA by chelating DNA base sites. If a metal ion at certain concentration binds mainly to DNA bases, it can decrease the binding constants between SPFX and DNA through competing with SPFX. While if a metal ion at certain concentration mainly binds to phosphate groups of DNA, it can increase the binding constants by building a bridge between SPFX and DNA. If a metal ion at certain concentrations binds neither to bases nor phosphate groups in DNA, it will have no effect on the binding constant between SPFX and DNA. Our result supports Palumbo's conclusion that the binding between SPFX and the phosphata groups is the precondition for the combination between SPFX and DNA, which is stabilized through stacking interactions between the condensed rings of SPFX and DNA bases.  相似文献   

6.
The rates of dissociation of 2 equiv of various metal ions [Ca(II), Cd(II), Pr(III), Nd(III), Sm(III), Eu(III), Gd(III), Tb(III), Dy(III), Ho(III), Er(III), Yb(III), and Lu(III)] from the primary CD and EF metal ion binding sites of parvalbumin (isotype pI = 4.75) from codfish (Gadus callarius L) were measured by stopped-flow techniques. The removal or replacement of metal ions was monitored by changes in sensitized Tb(III) luminescence or in intrinsic protein tryptophan fluorescence as quenching ions [Eu(III) or Yb(III)] were bound or removed or as the apoprotein was formed. In experiments wherein the bound metal ions were removed by mixing the parvalbumin with an excess of 1,2-diaminocyclohexanetetraacetic acid (DCTA), the kinetic traces were best fit by a double exponential with koff rate constants of 1.07 and 5.91 s-1 for Ca(II), 1.54 and 10.5 s-1 for Cd(II), and approximately 0.05 and approximately 0.5 s-1 for all of the trivalent lanthanide ions. In experiments wherein the bound metal ions were exchanged with an excess of a different metal ion, pseudo-first-order rate constants were proportional to the concentration of excess attacking metal ion for both the fast and slow processes in most experiments. In these cases, extrapolation of the rate constants to zero concentration of attacking metal ion gave values which agree well with the DCTA scavenging results. This finding demonstrates that the off rate constants do not depend on the occupancy of the neighboring site and therefore implies that there is no significant cooperativity in metal ion binding between the two sites in parvalbumin.  相似文献   

7.
The tripeptide glutathione (gamma-L-Glu-L-Cys-Gly, GSH) is thought to play an important role in the biological processing of antimony drugs. We have studied the complexation of the antileishmanial drug potassium antimony(III) tartrate to GSH in both aqueous solution and intact red blood cells by NMR spectroscopy and electrospray ionization mass spectrometry. The deprotonated thiol group of the cysteine residue is shown to be the only binding site for Sb(III), and a complex with the stoichiometry [Sb(GS)3] is formed. The stability constant for [Sb(GS)3] was determined to be log K 25 (I = 0.1 M, 298 K) based on a competition reaction between tartrate and GSH at different pH* values. In spite of being highly thermodynamically stable, the complex is kinetically labile. The rate of exchange of GSH between its free and Sb-bound form is pH-dependent, ranging from slow exchange on the 1H-NMR timescale at low pH (2 s-1 at pH 3.2) to relatively rapid exchange at biological pH (> 440 s-1). Such facile exchange may be important in the transport of Sb(III) in various biofluids and tissues in vivo. Our spin-echo 1H-NMR data show that Sb(III) rapidly entered red blood cell walls and was complexed by intracellular glutathione.  相似文献   

8.
The kinetics and thermodynamics of Al(III) exchange between aluminum citrate (AlL) and human serum transferrin were investigated in the 7.2-8.9 pH range. The C-site of human serum apotransferrin in interaction with bicarbonate removes Al(III) from Al citrate with an exchange equilibrium constant K1 = (2.0 +/- 0.6) x 10(-2); a direct second-order rate constant k1 = 45 +/- 3 M(-1) x s(-1); and a reverse second-order rate constant k(-1) = (2.3 +/- 0.5) x 10(3) M(-1) x s(-1). The newly formed aluminum-protein complex loses a single proton with proton dissociation constant K1a = (15 +/- 3) nM to yield a first kinetic intermediate. This intermediate then undergoes a modification in its conformation followed by two proton losses; first-order rate constant k2 = (4.20 +/- 0.02) x 10(-2) s(-1) to produce a second kinetic intermediate, which in turn undergoes a last slow modification in the conformation to yield the aluminum-loaded transferrin in its final state. This last process rate-controls Al(III) uptake by the N-site of the protein and is independent of the experimental parameters with a constant reciprocal relaxation time tau3(-1) = (6 +/- 1) x 10(-5) x s(-1). The affinities involved in aluminum uptake by serum transferrins are about 10 orders of magnitude lower than those involved in the uptake of iron. The interactions of iron-loaded transferrins with transferrin receptor 1 occur with average dissociation constants of 3 +/- 1 and 5 +/- 1 nM for the only C-site iron-loaded and of 6.0 +/- 0.6 and 7 +/- 0.5 nM for the iron-saturated ST in the absence or presence of CHAPS, respectively. No interaction is detected between receptor 1 and aluminum-saturated or mixed C-site iron-loaded/N-site aluminum-loaded transferrin under the same conditions. The fact that aluminum can be solubilized by serum transferrin in biological fluids does not necessarily imply that its transfer from the blood stream to cytoplasm follows the receptor-mediated pathway of iron transport by transferrins.  相似文献   

9.
The NMR diffusometry technique, based on the measurement of the diffusion coefficient of a ligand in the absence and in the presence of its macromolecular partner, was used to study the affinity for human serum albumin (HSA) of four gadolinium complexes, potential or already used magnetic resonance imaging contrast agents. Diamagnetic lanthanum(III) ion or europium(III) ion, which has the advantage of shifting the NMR signals far away from those of the macromolecule, was used to avoid the excessive broadening of the NMR signals induced by the gadolinium(III) ion. Titration experiments, in which the HSA concentration was kept constant and the concentration of the europium or lanthanum chelate was varied, were performed to evaluate the association constant and the number of binding sites. Some additional information about the kinetics of the exchange between the free and the bound chelate was also obtained. Competition experiments with ibuprofen and salicylate, which are ligands with a known affinity for the macromolecule and for which the binding site is known, were also performed to get information about the binding site of the contrast agents.  相似文献   

10.
The ductus deferens smooth muscle tumor cell line (DDT1MF-2) contains receptors for, and is stimulated by, androgens. Cells cultured in the absence of androgens maintain a basal level of androgen receptors. Following incubation with various concentrations of the synthetic androgen methyltrienolone (R1881) for 1-6 h, the concentration of these receptors increased from 6.0 to 12.2 fmol/micrograms of DNA, while the equilibrium dissociation constant (Kd) of 0.5 nM for this steroid remained unchanged. The steroid-induced increase in androgen receptor levels was specific for androgens and dependent upon protein synthesis. The mechanism of receptor augmentation was examined by utilization of isotopically dense amino acids to determine rates of receptor appearance and degradation in the presence or absence of [3H]R1881. In the absence of androgens, the half-life of the androgen receptor was 3.1 h, with a rate constant (kD) of 0.22/h. In the presence of 1 nM [3H]R1881, however, the half-life was 6.6 h, with kD = 0.11/h. The rate constant for receptor synthesis (ks) in the absence or presence of [3H]R1881 was calculated to be 1.35 and 2.23 fmol/micrograms of DNA/h, respectively. Thus, androgen-induced androgen-receptor augmentation is explained by an increase both in receptor half-life and in rates of receptor synthesis.  相似文献   

11.
2-Carboxylbenzaldehyde thiosemicarbazone (HL), and its three lanthanide (III) complexes, LnL(3) x 4H(2)O [Ln(III)=La, Sm, Eu], have been synthesized in water. The complexes were characterized by elemental analyses, molar conductivity and IR spectra. The crystal structure of [Sm(2)L(6)(CH(3)OH)(4)] x 7.5CH(3)OH x 0.5H(2)O obtained from methanol solution was determined by X-ray diffraction analysis, crystallized in the triclinic system, space group P-1, Z=1, a=12.217 (2)A, b=14.706 (2)A, c=15.035 (2)A, alpha=111.84(1) degrees , beta=103.47(1) degrees , gamma=104.24(1) degrees , R(1)=0.0290. It has symmetrical (mu-OCO)(2), (mu-O)(2) and disamarium(III) units. The coordination geometry of each Sm(III) ion is a distorted tetradecahedron with nine oxygen atoms. In addition, the DNA-binding properties of the ligand and its complexes have been investigated by absorption, fluorescence, and viscosity measurements. The experimental results indicate that the ligand and the Sm-complex can bind to DNA, but the other two complexes cannot; the binding affinity of the Sm-complex is higher than that of the ligand and the intrinsic binding constant K(b) of the complex is 3.22 x 10(5)M(-1).  相似文献   

12.
Salts and polyamines have a variety of effects on the physical properties of DNA, including stabilization against thermal melting. We wished to gain greater insight into the mechanism of this stabilization by ascertaining its effect on the dynamics of base opening and closing reactions, as measured by NMR. Since the binding of spermidine(3+) is influenced by salt, and since spermidine may act as a base catalyst in proton exchange reactions, we have undertaken a study of salt and base catalyst effects on the imino proton exchange kinetics of a model oligomeric DNA. The selective longitudinal NMR relaxation rates of the hydrogen-bonded imino protons of the self-complementary octadeoxyribonucleotide d(GGAATTCC) monitor the rate of the base-catalyzed chemical exchange of these protons with solvent water. The exchange rates thus obtained provide a sensitive measure of the base-pair opening reactions of the DNA duplex. Under conditions of low pH and no added base catalyst, the NMR relaxation rates allow the determination of kd, the rate constant for the dissociation of the octameric duplex into single strands. Titration with the base catalyst tris(hydroxymethyl)aminomethane allows the determination of kop, the rate constant for the localized opening of individual base pairs, prior to dissociation. A significant Na+ concentration dependence is found for kd. From an analysis of this dependence, it is determined that 0.6 +/- 0.1 sodium ion is released during the dissociation event. The activation energy for helix dissociation (200 +/- 5 kJ/mol) is not dependent on the sodium ion concentration, indicating that the dissociation is entropically driven by the release of bound sodium ions.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
The DNA polymerase activity of the near homogeneous, multisubunit DNA polymerase-primase from Drosophila melanogaster embryos has been compared to Escherichia coli DNA polymerase III core, DNA polymerase III, and DNA polymerase III holoenzyme. The rate of deoxynucleotide incorporation by the Drosophila polymerase on singly primed phi X174 DNA is similar to that observed with equivalent levels of DNA polymerase III holoenzyme in the absence of E. coli single-stranded DNA binding protein. However, analysis of the DNA products indicates that the Drosophila polymerase is less processive than DNA polymerase III holoenzyme, and closely resembles DNA polymerase III. The Drosophila polymerase-primase contains neither 3'-5' exonuclease nor RNase H-like activities, and catalyzes no significant pyrophosphate exchange. There is a low level of DNA-dependent ATPase activity which can be eliminated by a second glycerol gradient sedimentation (Kaguni, L.S., Rossignol, J.-M., Conaway, R.C., and Lehman, I.R. (1983) Proc. Natl. Acad. Sci. U. S. A. 80, 2221-2225). Although lacking a 3'-5' exonuclease, the replication fidelity of the D. melanogaster polymerase is similar to that of E. coli DNA polymerase III holoenzyme which possesses such an activity.  相似文献   

14.
We investigated whether a combination of recently introduced methods, total demineralization and ion‐exchange columns, would increase DNA recovery from old bone. Ten bone samples taken after a burial period of ∼60 years were used in this study. Bone powder was digested using total or incomplete demineralization. DNA was extracted by the standard organic method. The DNA extract was purified with ion‐exchange columns or QIAquick® spin columns. The efficiency of different DNA extraction methods was compared in terms of DNA concentration, inhibitors generated by real‐time PCR, and conventional STR typing results. The mean DNA concentration using the total demineralization method is ∼3 times higher than that using the incomplete demineralization method. For DNA purification, the method using QIAquick® spin columns appeared to yield approximately double the DNA than the method using ion‐exchange columns. Furthermore, 2 out of 10 samples showed higher levels of inhibition with CT values of IPC ≥30 cycles when using only ion‐exchange columns. In STR results, total demineralization yielded more locus profiles by 4.2 loci than incomplete demineralization, and QIAquick® spin columns also yielded more locus profiles by 3.5 loci than ion‐exchange columns. Total demineralization of bone powder significantly increased DNA yield and improved STR typing results. However, the use of ion‐exchange columns was not efficient when compared with the method using QIAquick® spin columns. It is suggested that the combination of total demineralization and QIAquick® spin columns lead to greatly improved STR typing results. Am J Phys Anthropol 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

15.
T E Strzelecka  R L Rill 《Biopolymers》1990,30(7-8):803-814
Aqueous solutions of DNA fragments with a contour length (500 A) near the persistence length at DNA concentrations ranging from 10 to 290 mg/mL solvent and a constant supporting electrolyte concentration of 0.01 M (predominantly NaCl) were examined by 23Na-nmr spectroscopy at temperatures of 20, 40, and 60 degrees C. Over the higher portion of this concentration range (greater than 100 mg/ml) the DNA solutions undergo a complex series of transitions between different anisotropic, liquid crystalline phases (T. E. Strzelecka and R. L. Rill, Biopolymers, in press). Counterions in solutions of strong polyelectrolytes are usually described in terms of a two-state model as free or "bound" (influenced by the electrostatic field of the polyanion). The longitudinal relaxation rate (R1 = 1/T1) at all DNA concentrations decreased with increasing temperature, demonstrating fast exchange between free and bound counterions. R1 increased nearly linearly with increasing DNA phosphate/sodium ratio in the isotropic domain until the onset of anisotropic phase formation, in agreement with similar nmr studies conducted at low DNA concentrations. The value of R1,b = 194 +/- 7 Hz obtained for the isotropic phase from 10 to 100 mg DNA/mL at 20 degrees C was in agreement with values reported previously. A nonlinear increase in R1 with DNA concentration was observed upon onset of anisotropic phase formation, indicating an increase in the product of the fraction of bond ions times their relaxation rate (r.R1,b). The spectral lineshape of all isotropic samples was Lorentzian. Spectra of anisotropic samples exhibited low magnitude quadrupole splitting of less than or equal to 400 Hz correlated with appearance of a cholesteric phase with pitch approximately 2 microns. The magnitude of the quadrupole splitting decreased with increasing DNA concentration at low temperatures and increased with concentration at high temperatures. At all concentrations the quadrupole splitting decreased then increased with temperature. These temperature- and concentration-dependent changes in quadrupole splitting are consistent with an angle between the DNA helix axis and the principal component (VZZ) of the local electric field gradient tensor near the "magic angle" of 54.7 degrees.  相似文献   

16.
J J Led  E Neesgaard 《Biochemistry》1987,26(1):183-192
A detailed analysis of the stability and activity of Mn(II) human carbonic anhydrase I and the kinetics and mechanism of its catalysis of the HCO3-/CO2 exchange have been performed at pH 8.5. The analysis was based on the paramagnetic relaxation rates R1p and R2p of the 13C atom of HCO3- in the Mn2+/apoenzyme/HCO3-/CO2 system and the HCO3(-)----CO2 interconversion rate obtained by the magnetization-transfer technique. The R1p and R2p rates were measured as functions of the temperature, magnetic field strength, and substrate and apoenzyme concentrations and were interpreted on the basis of the Solomon-Bloembergen-Morgan theories and general equations for the ligand exchange [Led, J. J., & Grant, D. M. (1977) J. Am. Chem. Soc. 99, 5845-5858]. From the analysis of the data, a formation constant for the Mn(II) enzyme of log KMAM = 5.8 +/- 0.4 was obtained while the activity of the Mn(II) enzyme, measured as the HCO3-/CO2 interconversion rate at [HCO3-] = 0.100 M and pH 8.5, was found to be about 4% of that of the native Zn(II) enzyme. However, an effective dissociation constant KeffHCO3- less than or approximately 12 mM and a maximal exchange rate constant kcatexch approximately equal to 400 s-1, also derived by the analysis, result in an apparent second-order rate constant kcatexch/KeffHCO3- only a factor of 4 smaller than the corresponding rate constant for the native Zn(II) isoenzyme I. Most conspicuously, the resulting distance of only 2.71 +/- 0.03 A between the Mn2+ ion of the enzyme and the 13C atom of HCO3- in the enzyme-bicarbonate complex indicates that the bicarbonate is bound to the metal ion by two of its oxygen atoms in the central catalytic step, thereby supporting the modified Zn(II)-OH mechanism [Lindskog, S., Engberg, P., Forsman, C., Ibrahim, S. A., Jonsson, B.-H., Simonsson, I., & Tibell, L. (1984) Ann. N.Y. Acad. Sci. 429, 61-75 (and references cited therein)]. In contrast, this binding mode differs from the structure of the complexes suggested in the rapid-equilibrium kinetic model [Pocker, Y., & Deits, T. L. (1983) J. Am. Chem. Soc. 105, 980-986; Pocker, Y., & Deits, T. L. (1984) Ann. N.Y. Acad. Sci. 429, 76-83].  相似文献   

17.
18.
Aluminum ion perturbs the activity of a number of physiologically important enzymes, including members of a family of guanine nucleotide-binding proteins (G-proteins). G-proteins couple cellular receptor proteins to a variety of effector enzymes (including adenylate cyclase, phospholipase C, and the rod photoreceptor phosphodiesterase). We show herein that subnanomolar concentrations of free aluminum ion, produced in a carefully defined and kinetically stable manner through the buffering of total aluminum at 0.1-1.0 mM with calculated ratios of chelating agents, inhibit both the receptor-mediated activation and the self-inactivating GTPase activity of the rod photoreceptor G-protein, Gv. In the presence of 4 X 10(-10) M free aluminum ion, GTPase activity is inhibited from about 25-60% as the magnesium ion concentration is reduced from 10(-3) to about 5 X 10(-5) M. The principal effect of aluminum ion upon Gv is to inhibit receptor catalyzed nucleotide exchange. Binding of the GTP analog 5'-guanylyl imidodiphosphate can be reduced by as much as 90% by aluminum ion following subsaturating rhodopsin stimulation. Aluminum ion can produce either competitive or mixed noncompetitive inhibition of rhodopsin-catalyzed Gv activation and GTPase activity, as a function of whether Gv undergoes single (competitive), or multiple (mixed noncompetitive) nucleotide exchanges. The rod photoreceptor phosphodiesterase is only slightly inhibited by similar aluminum ion activities. Light- and Gv-coupled phosphodiesterase activation exhibits both a lower maximum rate of cyclic guanosine monophosphate hydrolysis and a slower inactivation in the presence of aluminum ion activities from about 10(-12) - 10(-10) M. These data suggest that intracellular free aluminum ion concentrations in the subnanomolar range could markedly affect the ability of cells to transduce extracellular signals. Interestingly, the combination of Al3+ and F- to produce the fluoro-aluminate species (AlFx) also inhibits the GTPase of G-proteins, although the mechanism of inhibition (e.g. binding to the G-protein.Mg2+.GDP complex) is totally distinct from that observed for free Al3+ and the overall effect on signal transduction (e.g. enhanced signal amplification) is in complete opposition to that observed for free Al3+.  相似文献   

19.
Chicken ovotransferrin, an iron binding protein, has two metal binding sites (amino (N) and carboxy (C) terminal sites). It binds Cu(II), Al(III), Co(II), and other metals, as well as Fe(III). In this study, the selectivity and cooperativity of the N and C sites on Al(III), Co(II), and Tb(III) binding were investigated. Metals were classified into two groups according to their site preference. Co(II) and Al(III) bound to the N site more preferably than to the C site, whereas Tb(III) bound to the C site more preferably. On Fe(III) binding, the binding constant of Fe(III) becomes larger when the other site is already occupied. Thus, positive cooperativity is seen. In the present study, the binding cooperativities of Co(II), Tb(III), and Al(III) as to the N and C sites were investigated. On Co(II) and Tb(III) binding, no cooperativity was observed, as in the case of Cu(II) [Yamamura, T. et al. (1985) in Proteins of Iron Storage and Transport (Spik, G., Montreuil, J., Crichton, R.R., & Mazurier, J., eds.) pp. 53-56, Elsevier Science Publ. B.V., Amsterdam]. In contrast, negative cooperativity was observed on Al(III) binding. Based on a model proposed by Yamamura et al. [Yamamura, T. et al. (1985) ibid.], the ratio of the binding constants, KC/KN, and the stacking coefficient, Kst, were estimated. KC/KN is 2.2 +/- 0.4 for the Tb(III) ion, 0.5 +/- 0.1 for the Co(II) ion, and 0.12 +/- 0.02 for the Al(III) ion. Kst (= 1 in a non-cooperative case) is 0.98 +/- 0.02 for the Tb(III) ion, 1.03 +/- 0.02 for the Co(II) ion, and 0.55 +/- 0.22 for the Al(III) ion.  相似文献   

20.
Binding of E.coli lac repressor to non-operator DNA*   总被引:4,自引:2,他引:2       下载免费PDF全文
It is shown by melting profile analysis of lac repressor-DNA complexes that repressor binds tightly and preferentially (relative to single-stranded DNA) to double-stranded non-operator DNA. This binding stabilizes the DNA against melting and the repressor against thermal denaturation. Analysis of the extent of stabilization and the rate of dissociation of repressor from non-operator DNA as a function of sodium ion concentration shows, in confirmation of other studies,(3,4) that the binding constant (K(RD)) is very ionic strength dependent; K(RD) increases from approximately 10(6) M(-1) at approximately 0.1 M Na(+) to values in excess of 10(10) M(-1) at 0.002 M Na(+). Repressor bound to non-operator DNA is not further stabilized against thermal denaturation by inducer binding, indicating that the inducer and DNA binding sites probably represent separately stabilized local conformations. Transfer melting experiments are used to measure the rate of dissociation of repressor from operator DNA. These experiments show that most of the ionic strength dependence of the binding constant is in the dissociation process; the estimated dissociation rate constant decreases from greater than 10(-1) sec(-1) at [Na(+)] >/= 0.02 M to less than 10(-4) sec(-1) at [Na(+)] 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