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.
Regulation of the F-actin severing activity of gelsolin by Ca2+ has been investigated under physiologic ionic conditions. Tryptophan fluorescence intensity measurements indicate that gelsolin contains at least two Ca2+ binding sites with affinities of 2.5 x 10(7) M-1 and 1.5 x 10(5) M-1. At F-actin and gelsolin concentrations in the range of those found intracellularly, gelsolin is able to bind F-actin with half-maximum binding at 0.14 microM free Ca2+ concentration. Steady-state measurements of gelsolin-induced actin depolymerization suggest that half-maximum depolymerization occurs at approximately 0.4 microM free Ca2+ concentration. Dynamic light scattering measurements of the translational diffusion coefficient for actin filaments and nucleated polymerization assays for number concentration of actin filaments both indicate that severing of F-actin occurs slowly at micromolar free Ca2+ concentrations. The data suggest that binding of Ca2+ to the gelsolin-F-actin complex is the rate-limiting step for F-actin severing by gelsolin; this Ca2+ binding event is a committed step that results in a Ca2+ ion bound at a high-affinity, EGTA-resistant site. The very high affinity of gelsolin for the barbed end of an actin filament drives the binding reaction equilibrium toward completion under conditions where the reaction rate is slow.  相似文献   

2.
3.
Mechanism of the interaction of human platelet profilin with actin   总被引:24,自引:4,他引:20  
We have reexamined the interaction of purified platelet profilin with actin and present evidence that simple sequestration of actin monomers in a 1:1 complex with profilin cannot explain many of the effects of profilin on actin assembly. Three different methods to assess binding of profilin to actin show that the complex with platelet actin has a dissociation constant in the range of 1 to 5 microM. The value for muscle actin is similar. When bound to actin, profilin increases the rate constant for dissociation of ATP from actin by 1,000-fold and also increases the rate of dissociation of Ca2+ bound to actin. Kinetic simulation showed that the profilin exchanges between actin monomers on a subsecond time scale that allows it to catalyze nucleotide exchange. On the other hand, polymerization assays give disparate results that are inconsistent with the binding assays and each other: profilin has different effects on elongation at the two ends of actin filaments; profilin inhibits the elongation of platelet actin much more strongly than muscle actin; and simple formation of 1:1 complexes of actin with profilin cannot account for the strong inhibition of spontaneous polymerization. We suggest that the in vitro effects on actin polymerization may be explained by a complex mechanism that includes weak capping of filament ends and catalytic poisoning of nucleation. Although platelets contain only 1 profilin for every 5-10 actin molecules, these complex reactions may allow substoichiometric profilin to have an important influence on actin assembly. We also confirm the observation of I. Lassing and U. Lindberg (1985. Nature [Lond.] 318:472-474) that polyphosphoinositides inhibit the effects of profilin on actin polymerization, so lipid metabolism must also be taken into account when considering the functions of profilin in a cell.  相似文献   

4.
Understanding how amyloid-β peptide interacts with living cells on a molecular level is critical to development of targeted treatments for Alzheimer's disease. Evidence that oligomeric Aβ interacts with neuronal cell membranes has been provided, but the mechanism by which membrane binding occurs and the exact stoichiometry of the neurotoxic aggregates remain elusive. Physiologically relevant experimentation is hindered by the high Aβ concentrations required for most biochemical analyses, the metastable nature of Aβ aggregates, and the complex variety of Aβ species present under physiological conditions. Here we use single molecule microscopy to overcome these challenges, presenting direct optical evidence that small Aβ(1-40) oligomers bind to living neuroblastoma cells at physiological Aβ concentrations. Single particle fluorescence intensity measurements indicate that cell-bound Aβ species range in size from monomers to hexamers and greater, with the majority of bound oligomers falling in the dimer-to-tetramer range. Furthermore, while low-molecular weight oligomeric species do form in solution, the membrane-bound oligomer size distribution is shifted towards larger aggregates, indicating either that bound Aβ oligomers can rapidly increase in size or that these oligomers cluster at specific sites on the membrane. Calcium indicator studies demonstrate that small oligomer binding at physiological concentrations induces only mild, sporadic calcium leakage. These findings support the hypothesis that small oligomers are the primary Aβ species that interact with neurons at physiological concentrations.  相似文献   

5.
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   

6.
The distance separating the high-affinity binding sites of actin for a divalent metal ion and nucleotide was evaluated by using high-resolution proton NMR and EPR spectroscopy. Replacement of the Ca2+ or Mg2+ bound to the high-affinity divalent cation site of G-actin by trivalent lanthanide ions such as La3+, EU3+, or Gd3+ results in an increase in the mobility of the bound ATP as observed in the NMR spectra of G-actin monomers. Little difference was observed between the spectra obtained in the presence of the diamagnetic La3+ control and the paramagnetic ions Eu3+ and Gd3+ which respectively shift and broaden the proton resonances of amino acids in the vicinity of the binding site. Analysis of the NMR spectra indicates that the metal and nucleotide binding sites are separated by a distance of at least 16 A. In the past, the metal and ATP have been widely assumed to bind as a complex. Further verification that the two sites on actin are physically separated was obtained by using an ATP analogue with a nitroxide spin-label bound at the 6' position of the purine ring. An estimate of the distance was made between the site containing the ATP analogue and the paramagnetic ion, Mn2+, bound to the cation binding site. These EPR experiments were not affected by the state of polymerization of the actin. The data obtained by using this technique support the conclusion stated above, namely, that the cation and nucleotide sites on either G- or F-actin are well separated.  相似文献   

7.
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   

8.
Actin is a ubiquitous protein that is a major component of the cytoskeleton, playing an important role in muscle contraction and cell motility. At steady state, actin monomers and filaments (F-actin) coexist, and actin subunits continuously attach and detach at the filament ends. However, the size distribution of actin oligomers in F-actin solution has never been clarified. In this study, we investigated the size distribution of actin oligomers using photon-counting histograms. For this purpose, actin was labeled with a fluorescent dye, and the emitted photons were detected by confocal optics (the detection volume was of femtoliter (fL) order). Photon-counting histograms were analyzed to obtain the number distribution of actin oligomers in the detection area from their brightness, assuming that the brightness of an oligomer was proportional to the number of protomers. We found that the major populations at physiological ionic strength were 1-5mers. For data analysis, we successfully applied the theory of linear and helical aggregations of macromolecules. The model postulates three states of actin, i.e., monomers, linear polymers, and helical polymers. Here we obtained three parameters: the equilibrium constants for polymerization of linear polymers, K(l)=(5.2 +/- 1.1) x 10(6) M(-1), and helical polymers, K(h)=(1.6 +/- 0.5) x 10(7) M(-1); and the ratio of helical to linear trimers, gamma = (3.6 +/- 2.3) x 10(-2). The excess free energy of transforming a linear trimer to a helical trimer, which is assumed to be a nucleus for helical polymers, was calculated to be 2.0 kcal/mol. These analyses demonstrate that the oligomeric phase at steady state is predominantly composed of linear 1-5mers, and the transition from linear to helical polymers occurs on the level of 5-7mers.  相似文献   

9.
In this study, water proton relaxation rate (PRR) enhancements have been used to characterize the binding of metal ions to native ovalbumin, ovalbumin in which phosphate has been enzymatically cleaved from one or both of the two protein phosphoserines, and a heat-stabilized form of the protein (S-ovalbumin). With Scatchard plots constructed from water PRR enhancements, it was found that native ovalbumin and S-ovalbumin had one strong binding site for Mn2+ ion (KD approximately equal to 6.0 X 10(-4) M). Alkaline phosphatase treated ovalbumin, a protein having a single phosphoserine, had one Mn2+ binding site of slightly weaker affinity (KD approximately equal to 8.3 X 10(-4) M), while acid phosphatase treated ovalbumin, a dephosphorylated protein, had two much weaker Mn2+ ion binding sites (KD approximately equal to 1.3 X 10(-3) M). Competitive binding studies on the native protein suggested that Zn2+ ion competes with Mn2+ for the single strong-affinity site (KD approximately equal to 6.1 X 10(-3) M) while Mg2+ and Ca2+ do not. In a second set of experiments, the paramagnetic contribution to the 31P spin-lattice (T1P) and spin-spin (T2P) relaxation times at three separate magnetic field strengths was measured. Correlation times tau c characterizing Mn2+-31P dipolar relaxation were estimated from the ratios of T1P/T2P at a single field and from the ratios of spin-lattice relaxation rates at three different field strengths. The correlation times so obtained, ranging from about 0.7 to 7.7 ns at the three field strengths, were used in calculating distances from the bound Mn2+ ion to the phosphoserines of native ovalbumin, S-ovalbumin, and alkaline phosphatase treated ovalbumins. It was determined that the phosphate of phosphoserine-68 was 5.95 +/- 0.26 and 6.29 +/- 0.18 A from the Mn2+ in the native and alkaline phosphatase treated protein, respectively, and 6.99 +/- 0.30 A away from the Mn2+ in S-ovalbumin. The phosphate of phosphoserine-344 was determined to be 5.31 +/- 0.20 and 5.75 +/- 0.10 A from the Mn2+ ion in native ovalbumin and S-ovalbumin, respectively. The 13C nucleus of [1-13C]galactose enzymatically transferred to the nonreducing end of the ovalbumin oligosaccharide chain was not found to be significantly relaxed by Mn2+ bound to the protein, even at 1:1 stoichiometric ratio of metal:protein. Using this, we estimate the nonreducing terminal of the ovalbumin oligosaccharide to be at least 39 A from the metal ion binding site on the protein.  相似文献   

10.
Regulation of molluscan actomyosin ATPase activity   总被引:2,自引:0,他引:2  
The interaction of myosin and actin in many invertebrate muscles is mediated by the direct binding of Ca2+ to myosin, in contrast to modes of regulation in vertebrate skeletal and smooth muscles. Earlier work showed that the binding of skeletal muscle myosin subfragment 1 to the actin-troponin-tropomyosin complex in the presence of ATP is weakened by less than a factor of 2 by removal of Ca2+ although the maximum rate of ATP hydrolysis decreases by 96%. We have now studied the invertebrate type of regulation using heavy meromyosin (HMM) prepared from both the scallop Aequipecten irradians and the squid Loligo pealii. Binding of these HMMs to rabbit skeletal actin was determined by measuring the ATPase activity present in the supernatant after sedimenting acto-HMM in an ultracentrifuge. The HMM of both species bound to actin in the presence of ATP, even in the absence of Ca2+, although the binding constant in the absence of Ca2+ (4.3 X 10(3) M-1) was about 20% of that in the presence of Ca+ (2.2 X 10(4) M-1). Studies of the steady state ATPase activity of these HMMs as a function of actin concentration revealed that the major effect of removing Ca2+ was to decrease the maximum velocity, extrapolated to infinite actin concentration, by 80-85%. Furthermore, at high actin concentrations where most of the HMM was bound to actin, the rate of ATP hydrolysis remained inhibited in the absence of Ca+. Therefore, inhibition of the ATPase rate in the absence of Ca2+ cannot be due simply to an inhibition of the binding of HMM to actin; rather, Ca2+ must also directly alter the kinetics of ATP hydrolysis.  相似文献   

11.
We measured the concentration of calmodulin required to reverse inhibition by caldesmon of actin-activated myosin MgATPase activity, in a model smooth-muscle thin-filament system, reconstituted in vitro from purified vascular smooth-muscle actin, tropomyosin and caldesmon. At 37 degrees C in buffer containing 120 mM-KCl, 4 microM-Ca2+-calmodulin produced a half-maximal reversal of caldesmon inhibition, but more than 300 microM-Ca2+-calmodulin was necessary at 25 degrees C in buffer containing 60 mM-KCl. The binding affinity (K) of caldesmon for Ca2+-calmodulin was measured by a fluorescence-polarization method: K = 2.7 x 10(6) M-1 at 25 degrees C (60 mM-KCl); K = 1.4 x 10(6) M-1 at 37 degrees C in 70 mM-KCl-containing buffer; K = 0.35 x 10(6) M-1 at 37 degrees C in 120 mM-KCl- containing buffer (pH 7.0). At 37 degrees C/120 mM-KCl, but not at 25 degrees C/60 mM-KCl, Ca2+-calmodulin bound to caldesmon bound to actin-tropomyosin (K = 2.9 x 10(6) M-1). Ca2+ regulation in this system does not depend on a simple competition between Ca2+-calmodulin and actin for binding to caldesmon. Under conditions (37 degrees C/120 mM-KCl) where physiologically realistic concentrations of calmodulin can Ca2+-regulate synthetic thin filaments, Ca2+-calmodulin reverses caldesmon inhibition of actomyosin ATPase by forming a non-inhibited complex of Ca2+-calmodulin-caldesmon-(actin-tropomyosin).  相似文献   

12.
The binding sites of Mn2+, Co2+, and Gd3+ have been determined in triclinic lysozyme at pH 4.5 to 4.6. Mn2+ and Co2+ bind a site approximately 2.5 A from 1 of the oxygen atoms of the Glu-35 chain. The occupancy of the Mn2+ site is 0.22, corresponding to 1 bound ion for each 4.6 protein molecules. The occupancy of the Co2+ site is much lower, about 0.048. Gd3+ appears to be bound at two sites, the main one 2.5 A from an oxygen atom of the Glu-35 side chain, the other 3.1 A from an oxygen atom of the Asp-52 chain. The occupancy of both Gd3+ sites is low, 0.036 and 0.016, the latter being so low that the presence of the ion at this site is in doubt. The binding site of Mn2+ in the di(N-acetylglucosamine)-lysozyme complex has also been determined. It does not differ significantly from the Mn2+ binding site in the native protein, but the occupancy is lower, 0.16.  相似文献   

13.
The rate of association of actin with gelsolin was measured at various Ca2+ and ATP concentrations. The fraction of Ca2+-activated gelsolin was determined by quantitative evaluation of the association rates thereby assuming that Ca2+-binding gelsolin associates with actin and Ca2+-free gelsolin does not. A plot of the fraction of Ca2+-activated gelsolin vs. the free Ca2+ concentration revealed a sigmoidal shape suggesting that co-operative binding of Ca2+ ions is required for activation of gelsolin. A good fit of the experimental data by calculated binding curves was obtained if two Ca2+ ions were assumed to bind to actin in a highly co-operative manner. ATP decreased the rate of association of gelsolin with actin and bound to gelsolin at a low affinity (Kd = 32 microm for Ca2+-free and Kd = 400 microm for Ca2+-activated gelsolin). In contrast, a 1 : 1 gelsolin-actin complex was found to be activated for association with actin by a single Ca2+ ion in a non-co-operative manner.  相似文献   

14.
Gelsolin is a calcium binding protein that shortens actin filaments. This effect occurs in the presence but not in the absence of micromolar calcium ion concentrations and is partially reversed following removal of calcium ions. Once two actin molecules have bound to gelsolin in solutions containing Ca2+, one of the actins remains bound following chelation of calcium, so that the reversal of gelsolin's effect cannot be accounted for simply by its dissociation from the ends of the shortened filaments to allow for elongation. In this paper, the interactions with actin of the ethylene glycol bis(beta-aminoethyl ether)-N,N,N',N'-tetraacetic acid (EGTA) stable 1:1 gelsolin-actin complexes are compared with those of free gelsolin. The abilities of free or complexed gelsolin to sever actin filaments, nucleate filament assembly, bind to the fast growing (+) filament ends, and lower the filament size distribution in the presence of either Ca2+ or EGTA were examined. The results show that both free gelsolin and gelsolin-actin complexes are highly dependent on Ca2+ concentration when present in a molar ratio to actin less than 1:50. The gelsolin-actin complexes, however, differ from free gelsolin in that they have a higher affinity for (+) filament ends in EGTA and they cannot sever filaments in calcium. The limited reversal of actin-gelsolin binding following removal of calcium and the calcium sensitivity of nucleation by complexes suggest an alternative to reannealing of shortened filaments that involves redistribution of actin monomers and may account for the calcium-sensitive functional reversibility of the solation of actin by gelsolin.  相似文献   

15.
Caldesmon was purified to homogeneity from both chicken gizzard and bovine aortic smooth muscles. Caldesmon purified from bovine aorta was slightly larger than caldesmon purified from chicken gizzards (Mr = 140,000) when the two were compared electrophoretically. Caldesmon bound tightly to actin saturating at a molar ratio of 1 caldesmon monomer per 6.6 actin monomers. Ca2+-calmodulin appeared to reduce the affinity of caldesmon for actin. Caldesmon was also a potent inhibitor of heavy actomeromyosin ATPase activity producing a maximal effect at a ratio of 1 caldesmon monomer per 7-10 actin monomers. This effect was also antagonized by Ca2+-calmodulin. While caldesmon inhibited heavy actomeromyosin ATPase activity, it greatly enhanced binding of both unphosphorylated and phosphorylated heavy meromyosin to actin in the presence of MgATP, reducing the Kd for binding by a factor of 40 for each form of heavy meromyosin. Although we did identify a Ca2+-calmodulin-stimulated "caldesmon kinase" activity in caldesmon preparations purified under nondenaturing conditions, we observed no effect of phosphorylation (2 mol of PO4/mol of caldesmon) on the capacity to inhibit heavy actomeromyosin ATPase activity. Our results suggest that caldesmon could serve some role in smooth muscle function by enhancing cross-bridge affinity while inhibiting actomyosin ATPase activity.  相似文献   

16.
Arginase from Saccharomyces cerevisiae has long been known to be a metal ion-requiring enzyme as it requires heating at 45 degrees C in the presence of 10 mM Mn2+ for catalytic activation. Metals are also thought to play a structural role in the enzyme, but the identity of the structural metal and its precise structural role have not been defined. Analysis of the metal ions that bind to yeast arginase by atomic absorption spectroscopy reveals that there is a weakly associated Mn2+ that binds to the trimeric enzyme with a stoichiometry of 1.04 +/- 0.05 mol of Mn2+ bound per subunit and an apparent K'D value of 26 microM at pH 7.0 and 4 degrees C. A more tightly associated Zn2+ ion can only be removed by dialysis against chelating agents. In occasional preparations, this site contained some Mn2+; however, Zn2+ and Mn2+ together bind to high affinity sites with a stoichiometry of 1.14 +/- 0.25/mol of subunit. Both the loosely associated catalytic Mn2+ ion and the more tightly associated structural Zn2+ ion confer stability to the enzyme. Removal of the weakly bound Mn2+ ion results in a 3 degree C decrease in the midpoint of the thermal transition (T 1/2) (from 57 by 54 degrees C) as monitored by UV difference absorption spectroscopy. Removal of the tightly bound Zn2+ ion produces a 19 degrees C decrease in T 1/2 (to 38 degrees C). Similar results are obtained by circular dichroism measurements. When the Zn2+ ion is removed, the steady-state fluorescence intensity increases 100% as compared to the holoenzyme, with a shift in the emission maximum from 337 to 352 nm. This suggests that in the folded trimeric metalloenzyme, the tryptophan fluorescence is quenched and that upon removal of the structural metal, the quenching is relieved as tryptophan residues become exposed to more polar environments. Equilibrium sedimentation experiments performed after dialysis of the enzyme against EDTA demonstrate that arginase exists in a reversible monomer-trimer equilibrium, in the absence of metal ions, with a KD value of 5.05 x 10(-11) M2. In contrast, the native enzyme exists as a trimer with no evidence of dissociation when Mn2+ and Zn2+ are present (Eisenstein, E., Duong, L.T., Ornberg, R. L., Osborne, J.C., Jr., and Hensley, P. (1986) J. Biol. Chem. 261, 12814-12819). In summary, the study presented here demonstrates that binding of a weakly bound Mn2+ ion confers catalytic activity. In contrast, binding of a more tightly associated Zn2+ ion confers substantial stability to the tertiary and quaternary structure of the enzyme.  相似文献   

17.
The binding of specific ligand by linearly associating enzyme system M in equilibrium to M2 in equilibrium to M3 in equilibrium... has been discussed. It is assumed that ligand is bound in the region of the contact of monomers and free monomer retains the subsites for specific ligand binding. The character of the dependences of the amount of bound ligand on enzyme and ligand concentrations and the influence of specific ligand on the distribution between oligomeric enzyme forms have been analysed. The high concentrations of specific ligand provoke the dissociation of enzyme oligomers because of occupation of both subsites in monomeric form. The situation is discussed when a specific ligand is the substrate which is converted to the product in an intact binding site located in the region of the contact of monomers. The inhibitory effect of high substrate concentrations has been interpreted, taking into account the blocking of the association of inactive monomers into active oligomers.  相似文献   

18.
R K Gupta  R M Oesterling 《Biochemistry》1976,15(13):2881-2887
Rabbit muscle pyruvate kinase requires two divalent cations per active site for catalysis of the enolization of pyruvate in the presence of adenosine 5'-triphosphate (ATP). One divalent cation is bound directly to the enzyme and forms a second sphere complex with the bound ATP (site 1). The second divalent cation is directly coordinated to the phosphoryl groups of ATP and does not interact with the enzyme (site 2). The essential role of the divalent cation at site 1 is shown by the requirement for Mg2+ or Mn2+ for the enolization of pyruvate in the presence of the substitution inert Cr3+-ATP complex. The rate of detritiation of pyruvate shows a hyperbolic dependence of Mn2+ concentration in the presence of high concentrations of enzyme and Cr3+-ATP. A dissociation constant for Mn2+ from the pyruvate kinase-Mn2+-ATP-Cr3+-pyruvate complex of 1.3 +/- 0.5 muM is determined by the kinetics of detritiation of pyruvate and by parallel Mn2+ binding studies using electron paramagnetic resonance. The essential role of the divalent cation at site 2 is shown by the sigmoidal dependence of the rate of detritiation of pyruvate on Mn2+ concentration in the presence of high concentrations of enzyme and ATP yielding a dissociation constant of 29 +/- 9 muM for Mn2+ from site 2. This value is similar to the dissociation constant of the binary Mn-ATP complex (14 +/- 6 muM) determined under similar conditions. The rate of detritiation of pyruvate is proportional to the concentration of the pyruvate kinase-Mn2+-ATP-Mn2+-pyruvate complex, as determined by parellel kinetic and binding studies. Variation of the nature of the divalent cation at site 1 in the presence of CrATP causes only a twofold change in the rate of detritiation of pyruvate which does not correlate with the pKa of the metal-bound water. Variation of the nature of the divalent cation at both sites in the presence of ATP causes a sevenfold variation in the rate of detritiation or pyruvate that correlates with the pKa of the metal-bound water. The greater rate of enolization observed with CrATP fits this correlation, indicating that the electrophilicity of the nucleotide bound metal (at site 2) determines the rate of enolization of pyruvate.  相似文献   

19.
Apoconalbumin binds Mn(II) at two sites with association constants of K1 = 7 (+/- 1) X 10(4) and K2 = 0.4 (+/- 0.25) X 10(4) M-1. The binding is tighter in the presence of excess bicarbonate resulting in K1 = 1.8 (+/- 0.2) X 10(5) and K2 = 3 (+/- 2) X 10(4) M-1. The electron paramagnetic resonance spectrum (at both 9 and 35 GHz) of Mn(II) bound at the tight site reveals a rhombic distortion (lambda = E/D approximately equal to 0.25-0.31) in the protein ligand environment of the mental ion. An evaluation of the 1/pT1p, paramagnetic contribution to the longitudinal relaxation rate of solvent protons with Mn(II)-, Mn(III)-, and Fe(III)-derivatives of conalbumin revealed that the mental ion in each site of conalbumin is accessible to one water molecule. For Mn(II)-conalbumin and Mn(III)-conalbumin species, inner coordination sphere protons are rapidly exchanging with the bulk solvent, while slow exchange conditions prevail for Fe(III)-conalbumin.  相似文献   

20.
The effects of the divalent cations Mg2+, Mn2+ and Ca2+ on the Brownian rotational motion of fluorescently labeled myosin, heavy meromyosin and myosin subfragment-1 were measured by the method of time-resolved fluorescence depolarization. When Mg2+ was added to solutions of myosin or heavy meromyosin and EDTA, their rotational mobility increased. Ca2+ had no effect. Mn2+ increased the mobility of heavy meromyosin but decreased that of myosin. None of these divalent cations effected the mobility of subfragment-1. The binding of heavy meromyosin to actin was affected very little by Mg2+ or EDTA over a wide range of conditions. Divalent cations appear to change the swivel about which the heads of myosin rotate, presumably by binding to light chain 2 (also called DTNB light chain). However, the heads are still able to bind actin in nearly the same way whether Mg2+ is present or not. The concentration of free Mg2+ for the mid-point of the change in heavy meromyosin mobility is in good agreement with that for EDTA activation of ATPase activity. This suggests that EDTA activation is due to removal of Mg2+ bound to myosin itself.  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