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.
Henzl MT  Agah S  Larson JD 《Biochemistry》2004,43(34):10906-10917
Association of the parvalbumin AB and CD-EF domains was examined in Hepes-buffered saline, pH 7.4, employing fragments from rat alpha and beta. All of the interactions require Ca(2+). In saturating Ca(2+), the alpha AB/alpha CD-EF (alpha/alpha) complex displays an association constant of (7.6 +/- 0.4) x 10(7) M(-1). Ca(2+)-binding data for a mixture of the alpha fragments are compatible with an identical two-site model, yielding an average binding constant of (8.5 +/- 0.2) x 10(5) M(-1). The beta/beta interaction is significantly weaker, exhibiting an association constant of (3.0 +/- 0.6) x 10(6) M(-1). The Ca(2+)-binding constants for beta/beta are likewise diminished, at (1.0 +/- 0.1) x 10(5) and (2.3 +/- 0.2) x 10(4) M(-1). The magnitude of the apparent DeltaDeltaG(degree)' for Ca(2+) binding by alpha/alpha and beta/beta, at 3.4 kcal/mol, approaches that measured for the intact proteins (3.6 kcal/mol) and is substantially larger than the 1.5 kcal/mol value previously measured for the isolated CD-EF domains. This result suggests that the AB domain can modulate the Ca(2+) affinities of the CD and EF sites. Interestingly, the heterologous alpha/beta complex displays a larger association constant [(6.6 +/- 0.4) x 10(6) M(-1)] than the homologous beta/beta complex and heightened Ca(2+) affinity [binding constants of (1.3 +/- 0.1) x 10(6) and (8.8 +/- 0.2) x 10(4) M(-1)]. By contrast, beta/alpha associates more weakly than alpha/alpha and exhibits sharply reduced affinity for Ca(2+). Thus, the interaction between the beta AB domain and beta CD-EF domain may act to attenuate Ca(2+) affinity in the intact protein.  相似文献   

2.
Visinin-like protein 1 (VILIP-1) belongs to the neuronal calcium sensor family of Ca(2+)-myristoyl switch proteins that regulate signal transduction in the brain and retina. Here we analyze Ca(2+) and Mg(2+) binding, characterize metal-induced conformational changes, and determine structural effects of myristoylation and dimerization. Mg(2+) binds functionally to VILIP-1 at EF3 (ΔH = +1.8 kcal/mol and K(D) = 20 μM). Unmyristoylated VILIP-1 binds two Ca(2+) sequentially at EF2 and EF3 (K(EF3) = 0.1 μM and K(EF2) = 1-4 μM), whereas myristoylated VILIP-1 binds two Ca(2+) with lower affinity (K(D) = 1.2 μM) and positive cooperativity (Hill slope = 1.5). NMR assignments and structural analysis indicate that Ca(2+)-free VILIP-1 contains a sequestered myristoyl group like that of recoverin. NMR resonances of the attached myristate exhibit Ca(2+)-dependent chemical shifts and NOE patterns consistent with Ca(2+)-induced extrusion of the myristate. VILIP-1 forms a dimer in solution independent of Ca(2+) and myristoylation. The dimerization site is composed of residues in EF4 and the loop region between EF3 and EF4, confirmed by mutagenesis. We present the structure of the VILIP-1 dimer and a Ca(2+)-myristoyl switch to provide structural insights into Ca(2+)-induced trafficking of nicotinic acetylcholine receptors.  相似文献   

3.
Henzl MT  Larson JD  Agah S 《Biochemistry》2004,43(10):2747-2763
Rat alpha- and beta-parvalbumins have distinct monovalent cation-binding properties [Henzl et al. (2000) Biochemistry 39, 5859-5867]. Beta binds two Na(+) or one K(+), and alpha binds one Na(+) and no K(+). Ca(2+) abolishes these binding events, suggesting that the monovalent ions occupy the EF-hand motifs. This study compares alpha and beta divalent ion affinities in Na(+) and K(+) solutions. Solvent cation identity seriously affects alpha. In Hepes-buffered NaCl, at 5 degrees C, the macroscopic Ca(2+)-binding constants are 2.6 x 10(8) and 6.4 x 10(7) M(-1) and the Mg(2+) constants, 1.8 x 10(4) and 4.3 x 10(3) M(-1). In Hepes-buffered KCl, the Ca(2+) values increase to 2.9 x 10(9) and 6.6 x 10(8) M(-1) and the Mg(2+) values to 2.2 x 10(5) and 3.7 x 10(4) M(-1). Monte Carlo simulation of alpha binding data-employing site-specific constants and explicitly considering Na(+) binding-yields a K(Na) of 630 M(-1) and indicates that divalent ion-binding is positively cooperative. NMR data suggest that the lone Na(+) ion occupies the CD loop. Solvent cation identity has a smaller impact on beta. In Na(+), the Ca(2+) constants for the EF and CD sites are 2.3 x 10(7) and 1.5 x 10(6) M(-1), respectively; the Mg(2+) constants are 9.2 x 10(3) and 1.7 x 10(2) M(-1). In K(+), these values shift to 3.1 x 10(7) and 3.8 x 10(6) M(-1) and the latter to 1.4 x 10(4) and 2.9 x 10(2) M(-1). These data suggest that parvalbumin divalent ion affinity, particularly that of rat alpha, can be significantly attenuated by increased intracellular Na(+) levels.  相似文献   

4.
Agah S  Larson JD  Henzl MT 《Biochemistry》2003,42(37):10886-10895
Despite its higher net charge and reduced opportunities for favorable tertiary interactions, Ca(2+)-free rat beta-parvalbumin is more stable than rat alpha-parvalbumin. Under conditions wherein alpha denatures at 45.8 degrees C, beta denatures at 53.6 degrees. The homologous chicken beta isoform known as CPV3 also exhibits heightened stability-prompting an inquiry into the stabilizing influence of Pro-21 and Pro-26. Individual P21A and P26A mutations lower the T(m) of rat beta by 3.2 degrees, decreasing conformational stability by 0.74 kcal/mol. Simultaneous replacement of Pro-21 and Pro-26 essentially abolishes the excess stability (DeltaT(m) = -7.6 degrees; DeltaDeltaG(conf) = -1.77 kcal/mol). Significantly, the P21A/P26A variant displays Ca(2+) affinity virtually indistinguishable from wild-type beta, implying that structural alterations in the AB domain do not necessarily influence the divalent ion affinity of the CD-EF domain. The consequences of introducing proline at positions 21 and 26 in rat alpha were also examined. Whereas the H26P mutation raises the T(m) by 5.6 degrees (DeltaDeltaG(conf) = 1.25 kcal/mol), A21P lowers the T(m) by 8.5 degrees (DeltaDeltaG(conf) = -1.9 kcal/mol). Replacement of Ala-21 by proline in an alpha AB/beta CD-EF chimera increases the T(m) by 5.8 degrees (DeltaDeltaG(conf) = 0.95 kcal/mol), implying that the destabilization of alpha by Pro-21 results from steric conflict with a residue in the CD-EF domain. Consistent with that hypothesis, the K80S mutation markedly stabilizes alpha A21P, yielding a protein with a T(m) 2.0 degrees higher than wild-type alpha. The observed differences in stability resulting from proline addition/removal are largely consistent with alterations in main-chain and side-chain conformational entropy.  相似文献   

5.
S100B is a dimeric Ca(2+)-binding protein that undergoes a 90 +/- 3 degrees rotation of helix 3 in the typical EF-hand domain (EF2) upon the addition of calcium. The large reorientation of this helix is a prerequisite for the interaction between each subunit of S100B and target proteins such as the tumor suppressor protein, p53. In this study, Tb(3+) was used as a probe to examine how binding of a 22-residue peptide derived from the C-terminal regulatory domain of p53 affects the rate of Ca(2+) ion dissociation. In competition studies with Tb(3+), the dissociation rates of Ca(2+) (k(off)) from the EF2 domains of S100B in the absence and presence of the p53 peptide was determined to be 60 and 7 s(-)(1), respectively. These data are consistent with a previously reported result, which showed that that target peptide binding to S100B enhances its calcium-binding affinity [Rustandi et al. (1998) Biochemistry 37, 1951-1960]. The corresponding Ca(2+) association rate constants for S100B, k(on), for the EF2 domains in the absence and presence of the p53 peptide are 1.1 x 10(6) and 3.5 x 10(5) M(-)(1) s(-)(1), respectively. These two association rate constants are significantly below the diffusion control ( approximately 10(9) M(-)(1) s(-)(1)) and likely involve both Ca(2+) ion association and a Ca(2+)-dependent structural rearrangement, which is slightly different when the target peptide is present. EF-hand calcium-binding mutants of S100B were engineered at the -Z position (EF-hand 1, E31A; EF-hand 2, E72A; both EF-hands, E31A + E72A) and examined to further understand how specific residues contribute to calcium binding in S100B in the absence and presence of the p53 peptide.  相似文献   

6.
Relative to other parvalbumin isoforms, the mammalian beta-parvalbumin (oncomodulin) displays attenuated divalent ion affinity. High-resolution structural data for the Ca(2+)-bound protein have provided little insight into the physical basis for this behavior, prompting an examination of the unliganded state. This article describes the solution structure and peptide backbone dynamics of Ca(2+)-free rat beta-parvalbumin (beta-PV). Ca(2+) removal evidently provokes significant structural alterations. Interaction between the D helix and the AB domain in the Ca(2+)-bound protein is greatly diminished in the apo-form, permitting the D helix to straighten. There is also a significant reorganization of the hydrophobic core and a concomitant remodeling of the interface between the AB and CD-EF domains. These modifications perturb the orientation of the C and D helices, and the energetic penalty associated with their reversal could contribute to the low-affinity signature of the CD site. By contrast, Ca(2+) removal causes a comparatively minor perturbation of the E and F helices, consistent with the more typical divalent ion affinity observed for the EF site. Ca(2+)-free rat beta-PV retains structural rigidity on the picosecond-nanosecond timescale. At 20 degrees C, the majority of amide vectors show no evidence for motion on timescales above 20 ps, and the average order parameter for the entire molecule is 0.92.  相似文献   

7.
Henzl MT  Agah S 《Proteins》2006,62(1):270-278
Birds express three parvalbumins, one alpha isoform and two beta isoforms. The latter are known as avian thymic hormone (ATH) and avian parvalbumin 3. Although both were discovered in thymus tissue, and presumably function in T-cell maturation, they have been detected in other tissue settings. We have conducted detailed Ca2+- and Mg2+-binding studies on recombinant ATH and the C72S variant of CPV3, employing global analysis of isothermal titration calorimetry data. In Hepes-buffered saline, ATH binds Ca2+ with apparent microscopic binding constants of 2.4 +/- 0.2 x 10(8) and 1.0 +/- 0.1 x 10(8) M(-1). The corresponding values for CPV3-C72S are substantially lower, 4.5 +/- 0.5 x 10(7) and 2.4 +/- 0.2 x 10(7) M(-1), a 1.9-kcal/mol difference in binding free energy. Thus, the beta-parvalbumin lineage displays a spectrum of Ca2+-binding affinity, with ATH and the mammalian beta isoform at the high- and low-affinity extremes and CPV3 in the middle. Interestingly, despite its decreased Ca2+ affinity, CPV3-C72S exhibits increased affinity for Mg2+, relative to ATH. Whereas the latter displays Mg2+-binding constants of 2.2 +/- 0.2 x 10(4) and 1.2 +/- 0.1 x 10(4) M(-1), CPV3-C72S yields values of 5.0 +/- 0.8 x 10(4) and 2.1 +/- 0.3 x 10(4) M(-1).  相似文献   

8.
We have used the highly selective alpha(4)beta(1) inhibitor 2S-[(1-benzenesulfonyl-pyrrolidine-2S-carbonyl)-amino]-4-[4-methyl-2S-(methyl-[2-[4-(3-o-tolyl-ureido)-phenyl]-acetyl]-amino)-pentanoylamino]-butyric acid (BIO7662) as a model ligand to study alpha(4)beta(1) integrin-ligand interactions on Jurkat cells. Binding of [(35)S]BIO7662 to Jurkat cells was dependent on the presence of divalent cations and could be blocked by treatment with an excess of unlabeled inhibitor or with EDTA. K(D) values for the binding of BIO7662 to Mn(2+)-activated alpha(4)beta(1) and to the nonactivated state of the integrin that exists in 1 mm Mg(2+), 1 mm Ca(2+) were <10 pm, indicating that it has a high affinity for both activated and nonactivated integrin. No binding was observed on alpha(4)beta(1) negative cells. Through an analysis of the metal ion dependences of ligand binding, several unexpected findings about alpha(4)beta(1) function were made. First, we observed that Ca(2+) binding to alpha(4)beta(1) was stimulated by the addition of BIO7662. From solution binding studies on purified alpha(4)beta(1), two types of Ca(2+)-binding sites were identified, one dependent upon and the other independent of BIO7662 binding. Second, we observed that the metal ion dependence of ligand binding was affected by the affinity of the ligand for alpha(4)beta(1). ED(50) values for the metal ion dependence of the binding of BIO7762 and the binding of a lower affinity ligand, BIO1211, differed by 2-fold for Mn(2+), 30-fold for Mg(2+), and >1000-fold for Ca(2+). Low Ca(2+) (ED(50) = 5-10 microm) stimulated the binding of BIO7662 to alpha(4)beta(1). The effects of microm Ca(2+) closely resembled the effects of Mn(2+) on alpha(4)beta(1) function. Third, we observed that the rate of BIO7662 binding was dependent on the metal ion concentration and that the ED(50) for the metal ion dependence of BIO7662 binding was affected by the concentration of the BIO7662. These studies point to an even more complex interplay between metal ion and ligand binding than previously appreciated and provide evidence for a three-component coupled equilibrium model for metal ion-dependent binding of ligands to alpha(4)beta(1).  相似文献   

9.
Apparent Ca(2+)-binding constant (K(app)) of Caenorhabditis elegans troponin C (CeTnC) was determined by a fluorescence titration method. The K(app) of the N-domain Ca(2+)-binding site of CeTnC was 7.9+/-1.6 x 10(5) M(-1) and that of the C-domain site was 1.2+/-0.6 x 10(6) M(-1), respectively. Mg(2+)-dependence of the K(app) showed that both Ca(2+)-binding sites did not bind competitively Mg(2+). The Ca(2+) dissociation rate constant (k(off)) of CeTnC was determined by the fluorescence stopped-flow method. The k(off) of the N-domain Ca(2+)-binding site of CeTnC was 703+/-208 s(-1) and that of the C-domain site was 286+/-33 s(-1), respectively. From these values we could calculate the Ca(2+)-binding rate constant (k(on)) as to be 5.6+/-2.8 x 10(8) M(-1) s(-1) for the N-domain site and 3.4+/-2.1 x 10(8) M(-1) s(-1) for the C-domain site, respectively. These results mean that all Ca(2+)-binding sites of CeTnC are low affinity, fast dissociating and Ca(2+)-specific sites. Evolutional function of TnC between vertebrate and invertebrate and biological functions of wild type and mutant CeTnCs are discussed.  相似文献   

10.
Integrin receptor alpha(2)beta(1) requires micromolar Ca(2+) to bind to collagen and to the peptide GPC(GPP)(5)GFOGER(GPP)(5)GPC (denoted GFOGER-GPP, where O represents hydroxyproline), which contains the minimum recognition sequence for the collagen-binding alpha(2) I-domain (Knight, C. G., Morton, L. F., Peachey, A. R., Tuckwell, D. S., Farndale, R. W., and Barnes, M. J. (2000) J. Biol. Chem. 275, 35-40). Platelet adhesion to these ligands is completely dependent on alpha(2)beta(1) in the presence of 2 mm Mg(2+). However, we show here that this interaction was abolished in the presence of 25 microm EGTA. Adhesion of Glanzmann's thrombasthenic platelets, which lack the fibrinogen receptor alpha(IIb)beta(3), was also inhibited by micromolar EGTA. Mg(2+)-dependent adhesion of platelets was restored by the addition of 10 microm Ca(2+), but millimolar Ca(2+) was inhibitory. Binding of isolated alpha(2)beta(1) to GFOGER-GPP was 70% inhibited by 50 microm EGTA but, as with intact platelets, was fully restored by the addition of micromolar Ca(2+). 2 mm Ca(2+) did not inhibit binding of isolated alpha(2)beta(1) to collagen or to GFOGER-GPP. Binding of recombinant alpha(2) I-domain was not inhibited by EGTA, nor did millimolar Ca(2+) inhibit binding. Our data suggest that high affinity Ca(2+) binding to alpha(2)beta(1), outside the I-domain, is essential for adhesion to collagen. This is the first demonstration of a Ca(2+) requirement in alpha(2)beta(1) function.  相似文献   

11.
The independently folding C2 domain motif serves as a Ca(2+)-dependent membrane docking trigger in a large number of Ca(2+) signaling pathways. A comparison was initiated between three closely related C2 domains from the conventional protein kinase C subfamily (cPKC, isoforms alpha, beta, and gamma). The results reveal that these C2 domain isoforms exhibit some similarities but are specialized in important ways, including different Ca(2+) stoichiometries. In the absence of membranes, Ca(2+) affinities of the isolated C2 domains are similar (2-fold difference) while Hill coefficients reveal cooperative Ca(2+) binding for the PKC beta C2 domain but not for the PKC alpha or PKC gamma C2 domain (H = 2.3 +/- 0.1 for PKC beta, 0.9 +/- 0.1 for PKC alpha, and 0.9 +/- 0.1 for PKC gamma). When phosphatidylserine-containing membranes are present, Ca(2+) affinities range from the sub-micromolar to the micromolar (7-fold difference) ([Ca(2+)](1/2) = 0.7 +/- 0.1 microM for PKC gamma, 1.4 +/- 0.1 microM for PKC alpha, and 5.0 +/- 0.2 microM for PKC beta), and cooperative Ca(2+) binding is observed for all three C2 domains (Hill coefficients equal 1.8 +/- 0.1 for PKC beta, 1.3 +/- 0.1 for PKC alpha, and 1.4 +/- 0.1 for PKC gamma). The large effects of membranes are consistent with a coupled Ca(2+) and membrane binding equilibrium, and with a direct role of the phospholipid in stabilizing bound Ca(2+). The net negative charge of the phospholipid is more important to membrane affinity than its headgroup structure, although a slight preference for phosphatidylserine is observed over other anionic phospholipids. The Ca(2+) stoichiometries of the membrane-bound C2 domains are detectably different. PKC beta and PKC gamma each bind three Ca(2+) ions in the membrane-associated state; membrane-bound PKC alpha binds two Ca(2+) ions, and a third binds weakly or not at all under physiological conditions. Overall, the results indicate that conventional PKC C2 domains first bind a subset of the final Ca(2+) ions in solution, and then associate weakly with the membrane and bind additional Ca(2+) ions to yield a stronger membrane interaction in the fully assembled tertiary complex. The full complement of Ca(2+) ions is needed for tight binding to the membrane. Thus, even though the three C2 domains are 64% identical, differences in Ca(2+) affinity, stoichiometry, and cooperativity are observed, demonstrating that these closely related C2 domains are specialized for their individual functions and contexts.  相似文献   

12.
Brain spectrin alpha and beta chains bind 45Ca2+, as shown by the calcium overlay method. Flow dialysis measurements revealed eight high affinity binding sites/tetramer that comprise two binding components (determined by nonlinear regression analysis). The first component has one or two sites (kd = 2-30 x 10(-8) M), depending on the ionic strength of the binding buffer, with the remaining high affinity sites in the second component (kd = 1-3 x 10(-6) M). In addition, there is a variable, low affinity binding component (n = 100-400, kd = 1-2 x 10(-4) M). Magnesium inhibits calcium binding to the low affinity sites with a K1 = 1.21 mM. Proteolytic fragments from trypsin or chymotrypsin digests of brain spectrin bind 45Ca2+ if they include alpha domain IV, alpha domain III, or the amino-terminal half of the beta chain (but more than 25 kDa from the amino-terminal). These data suggest that calcium ions bind with high affinity to the putative EF-hands in alpha domain IV and to one site in the amino-terminal half of the beta chain that is associated with alpha domain IV in the native dimer. The localization is consistent with a direct calcium modulation of the spectrin-actin-protein 4.1 interaction. In addition, there appears to be one high affinity site near the hypersensitive region of alpha brain spectrin. All four proposed binding sites occur near probable calmodulin-binding or calcium-dependent protease cleavage sites.  相似文献   

13.
The use of competitive isothermal titration calorimetry (ITC) to measure high-affinity binding constants has been largely restricted to systems with a single binding site or multiple identical sites. This study demonstrates the extension of this approach to proteins with two nonequivalent EF-hand Ca(2+)-binding sites--rat beta parvalbumin and the S55D/E59D variant of rat alpha parvalbumin. The method involves simultaneous (global) least-squares analysis of titrations with Ca(2+), with Mg(2+), with Ca(2+) in the presence of Mg(2+), and with Ca(2+) or Mg(2+) in the presence of a competitive chelator (EDTA or EGTA). The Ca(2+) and Mg(2+) binding constants obtained for rat beta agree well with estimates obtained by flow dialysis. Although the Ca(2+) affinity of alpha S55D/E59D is too high to measure by flow dialysis, it was amenable to analysis using the ITC-based approach. The combined S55D and E59D mutations increase the Ca(2+) and Mg(2+) affinities of the mutated binding site by factors of 14 and 26, respectively. This behavior is consistent with that seen previously for the rat beta S55D variant.  相似文献   

14.
The effects of ryanodine, a selective inhibitor of the Ca(2+)-induced Ca2+ release mechanism, on caffeine-evoked changes in cytosolic Ca2+ concentration ([Ca2+]i) and catecholamine secretion were investigated using cultured bovine adrenal chromaffin cells. Caffeine (5-40 mM) caused a concentration-dependent transient rise in [Ca2+]i and catecholamine secretion in Ca2+/Mg(2+)-free medium containing 0.2 mM EGTA. Ryanodine (5 x 10(-5) M) alone had no effect on either [Ca2+]i or catecholamine secretion. Although the application of ryanodine plus caffeine caused the same increase in both [Ca2+]i and catecholamine secretion as those induced by caffeine alone, ryanodine (4 x 10(-7) - 5 x 10(-5) M) irreversibly prevented the increase in both [Ca2+]i and catecholamine secretion resulting from subsequent caffeine application over a range of concentrations. The secretory response to caffeine was markedly enhanced by replacement of Na+ with sucrose in Ca2+/Mg(2+)-free medium, and this enhanced response was also blocked by ryanodine. Caffeine was found to decrease the susceptibility of the secretory apparatus to Ca2+ in digitonin-permeabilized cells. These results indicate that caffeine mobilizes Ca2+ from intracellular stores, the function of which is irreversibly blocked by ryanodine, resulting in the increase in catecholamine secretion in the bovine adrenal chromaffin cell.  相似文献   

15.
Henzl MT  Larson JD  Agah S 《Biochemistry》2000,39(19):5859-5867
The mammalian genome encodes both alpha- and beta-parvalbumin isoforms. The rat beta-parvalbumin (aka "oncomodulin") is more stable than the alpha isoform at physiological pH and ionic strength, despite its substantially higher charge density and truncated C-terminal helix [Henzl, M. T., and Graham, J. S. (1999) FEBS Lett. 442, 241-245]. Reasoning that solvent interactions could contribute to this unexpected finding, we have examined the stabilities of the Ca(2+)-free alpha- and beta-parvalbumins as a function of Na(+) and K(+) concentration. Differential scanning calorimetry data suggest that, at physiological pH and ionic strength, the beta isoform binds roughly 2 equiv of Na(+) or a single equivalent of K(+) with moderate affinity. Under comparable conditions, the alpha isoform apparently binds just 1 equiv of Na(+) and essentially no K(+). Isothermal titration calorimetry experiments suggest that the bound monovalent ions occupy the EF-hand motifs. In 0.15 M K(+), at pH 7.4, the stability of the apo-beta-parvalbumin exceeds that of the alpha isoform by approximately 2.6 kcal/mol at 37 degrees C and by approximately 3.0 kcal/mol at 25 degrees C. The latter value represents a substantial fraction of the difference in Ca(2+)-binding free energies measured in vitro for the two proteins. Significantly, however, these results do not completely explain the paradoxical stability of the beta isoform, which maintains its higher melting temperature under all conditions examined.  相似文献   

16.
A phosphatase specific for the hydrolysis of 3-deoxy-d-manno-octulosonate (KDO)-8-phosphate was purified approximately 400-fold from crude extracts of Escherichia coli B. The hydrolysis of KDO-8-phosphate to KDO and inorganic phosphate in crude extracts of E. coli B, grown in phosphate-containing minimal medium, could be accounted for by the enzymatic activity of this specific phosphatase. No other sugar phosphate tested was an alternate substrate or inhibitor of the purified enzyme. KDO-8-phosphate phosphatase was stimulated three- to fourfold by the addition of 1.0 mM Co(+) or Mg(2+) and to a lesser extent by 1.0 mM Ba(2+), Zn(2+), and Mn(2+). The activity was inhibited by the addition of 1.0 mM ethylenediaminetetraacetic acid, Cu(2+), Ca(2+), Cd(2+), Hg(2+), and chloride ions (50% at 0.1 M). The pH optimum was determined to be 5.5 to 6.5 in both tris(hydroxymethyl)aminomethane-acetate and HEPES (N-2-hydroxyethylpiperazine-N'-2-ethanesulfonic acid) buffer. This specific phosphatase had an isoelectric point of 4.7 to 4.8 and a molecular weight of 80,000 +/- 6,000 as determined by molecular sieving and Ferguson analysis. The enzyme appeared to be composed of two identical subunits of 40,000 to 43,000 molecular weight. The apparent K(m) for KDO-8-phosphate was determined to be 5.8 +/- 0.9 x 10(-5) M in the presence of 1.0 mM Co(2+), 9.1 +/- 1 x 10(-5) M in the presence of 1.0 mM Mg(2+), and 1.0 +/- 0.2 x 10(-4) M in the absence of added Co(2+) or Mg(2+).  相似文献   

17.
In enteric synaptosomes of the rat, the role of voltage-dependent Ca(2+) channels in K(+)-induced VIP release and nitric oxide (NO) synthesis was investigated. Basal VIP release was 39 +/- 4 pg/mg, and cofactor-substituted NO synthase activity was 7.0 +/- 0.8 fmol. mg(-1). min(-1). K(+) depolarization (65 mM) stimulated VIP release Ca(2+) dependently (basal, 100%; K(+), 172.2 +/- 16.2%; P < 0.05, n = 5). K(+)-stimulated VIP release was reduced by blockers of the P-type (omega-agatoxin-IVA, 3 x 10(-8) M) and N-type (omega-conotoxin-GVIA, 10(-6) M) Ca(2+) channels by ~50 and 25%, respectively, but not by blockers of the L-type (isradipine, 10(-8) M), Q-type (omega-conotoxin-MVIIC, 10(-6) M), or T-type (Ni(2+), 10(-6) M) Ca(2+) channels. In contrast, NO synthesis was suppressed by omega-agatoxin-IVA, omega-conotoxin-GVIA, and isradipine by ~79, 70, and 70%, respectively, whereas Ni(2+) and omega-conotoxin-MVIIC had no effect. These findings are suggestive of a coupling of depolarization-induced VIP release primarily to the P- and N-type Ca(2+) channels, whereas NO synthesis is presumably dependent on Ca(2+) influx not only via the P- and N- but also via the L-type Ca(2+) channel. In contrast, none of the Ca(2+) channel blockers affected VIP release evoked by exogenous NO, suggesting that NO induces VIP secretion by a different mechanism, presumably involving intracellular Ca(2+) stores.  相似文献   

18.
Lee YH  Tanner JJ  Larson JD  Henzl MT 《Biochemistry》2004,43(31):10008-10017
In model peptide systems, Ca2+ affinity is maximized in EF-hand motifs containing four carboxylates positioned on the +x and -x and +z and -z axes; introduction of a fifth carboxylate ligand reduces the affinity. However, in rat beta-parvalbumin, replacement of Ser-55 with aspartate heightens divalent ion affinity [Henzl, M. T., et al. (1996) Biochemistry 35, 5856-5869]. The corresponding alpha-parvalbumin variant (S55D/E59D) likewise exhibits elevated affinity [Henzl, M. T., et al. (2003) Anal. Biochem. 319, 216-233]. To determine whether these mutations produce a variation on the archetypal EF-hand coordination scheme, we have obtained high-resolution X-ray crystallographic data for alpha S55D/E59D. As anticipated, the aspartyl carboxylate replaces the serine hydroxyl at the +z coordination position. Interestingly, the Asp-59 carboxylate abandons the role it plays as an outer sphere ligand in wild-type rat beta, rotating away from the Ca2+ and, instead, forming a hydrogen bond with the amide of Glu-62. Superficially, the coordination sphere in the CD site of alpha S55D/E59D resembles that in the EF site. However, the orientation of the Asp-59 side chain is predicted to stabilize the D-helix, which may contribute to the heightened divalent ion affinity. DSC data indicate that the alpha S55D/E59D variant retains the capacity to bind 1 equiv of Na+. Consistent with this finding, when binding measurements are conducted in K(+)-containing buffer, divalent ion affinity is markedly higher. In 0.15 M KCl and 0.025 M Hepes-KOH (pH 7.4) at 5 degrees C, the macroscopic Ca2+ binding constants are 1.8 x 10(10) and 2.0 x 10(9) M(-1). The corresponding Mg2+ binding constants are 2.7 x 10(6) and 1.2 x 10(5) M(-1).  相似文献   

19.
Ca(2+)-binding guanylyl cyclase-activating proteins (GCAPs) stimulate photoreceptor membrane guanylyl cyclase (retGC) in the light when the free Ca(2+) concentrations in photoreceptors decrease from 600 to 50 nM. RetGC activated by GCAPs exhibits tight dimerization revealed by chemical cross-linking (Yu, H., Olshevskaya, E., Duda, T., Seno, K., Hayashi, F., Sharma, R. K., Dizhoor, A. M., and Yamazaki, A. (1999) J. Biol. Chem. 274, 15547-15555). We have found that the Ca(2+)-loaded GCAP-2 monomer undergoes reversible dimerization upon dissociation of Ca(2+). The ability of GCAP-2 and its several mutants to activate retGC in vitro correlates with their ability to dimerize at low free Ca(2+) concentrations. A constitutively active GCAP-2 mutant E80Q/E116Q/D158N that stimulates retGC regardless of the free Ca(2+) concentrations forms dimers both in the absence and in the presence of Ca(2+). Several GCAP-2/neurocalcin chimera proteins that cannot efficiently activate retGC in low Ca(2+) concentrations are also unable to dimerize in the absence of Ca(2+). Additional mutation that restores normal activity of the GCAP-2 chimera mutant also restores its ability to dimerize in the absence of Ca(2+). These results suggest that dimerization of GCAP-2 can be a part of the mechanism by which GCAP-2 regulates the photoreceptor guanylyl cyclase. The Ca(2+)-free GCAP-1 is also capable of dimerization in the absence of Ca(2+), but unlike GCAP-2, dimerization of GCAP-1 is resistant to the presence of Ca(2+).  相似文献   

20.
In human heart failure, disturbances in Ca2+ homeostasis are well known but the fate of the Na,K-ATPase isoforms (alpha1beta1, alpha2beta1 and alpha3beta1), the receptors for cardiac glycosides, still remains under study. Microsomes have been purified from non-failing human hearts. As judged by the sensitivities of Na,K-ATPase activity to ouabain (IC50 values: 7.0 +/- 2.5 and 81 +/- 11 nM), 3H-ouabain-binding measurements at equilibrium with and without 10 mM K+ and by a biphasic ouabain dissociation process, at least two finctionally active Na,K-ATPase isozymes coexist in normal human hearts. These are demonstrated as a very high- and a high affinity ouabain-binding site. The KD values are 3.6 +/- 1.6 nM and 17 +/- 6 nM, respectively. The two dissociation rate constants are 42 x 10(4) min(-1) and 360 x 10(-4) min(-1). Addition of 10 mM K+ ions shifted the respective KD values for ouabain from 3.6 +/- 1.6 to 20 +/- 5 nM and from 17 +/- 6 nM to 125 +/- 25 nM, respectively. The isozymes involved are identified by comparing these three pharmacological parameters to those of each alpha/beta-isozyme separately expressed in Xenopus oocytes (9). In human heart, the very high affinity site for ouabain is the alpha1beta1 dimer and the high affinity site is alpha2beta1.  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