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 234 毫秒
1.
Binding between the microtubule-associated tau protein and S100b protein was demonstrated by affinity chromatography and cross-linking experiments and was manifested in the effect of S100b on tau protein phosphorylation by protein kinase II. All three expressions of the binding showed that S100b discriminates among the four species of tau, revealing for the first time that the different kinds of tau may differ functionally. Noncovalent interaction between tau and S100b depended on the presence of Ca2+ or Zn2+ and resulted in total inhibition of tau phosphorylation by protein kinase II. In the absence of reducing agent, covalent binding studies between Cys84 beta in the carboxyl-terminal region of the S100b-beta subunit and tau proteins confirmed interactions between the two proteins. It is suggested that the homologous calcium-binding domain that characterizes the carboxyl terminus of S100 and the tubulin subunit may be responsible for the common interaction of both proteins with tau proteins. The physicochemical relationship between S100 subunits and p11, the subunit of a substrate for tyrosine kinase, and their similarity in interaction with protein kinase substrates are discussed.  相似文献   

2.
To gauge similarities between S100b protein and calmodulin, interactions were observed between S100b and melittin and between S100b and tau, the microtubule-associated proteins. The interaction of melittin with S100b protein in the presence and absence of calcium was studied by fluorescence polarization, UV difference spectroscopy, and sulfhydryl derivatization. Whether calcium was present or not in the solution, melittin and S100b form a complex of molar ratios up to 2:1. Further binding of melittin occurred, but it resulted in precipitation of S100b, as is true of the corresponding case of melittin binding to calmodulin. In the absence of calcium, the interaction of melittin and S100b shielded the tryptophan (Trp) of the former protein and exposed cysteine-84 beta (Cys-84 beta) of the latter protein, leaving the tyrosine-16 beta (Tyr-16 beta) of S100b unaffected. Calcium addition to the complex partially restored the exposure of Trp of melittin and caused changes in the environment of Tyr-16 beta (unlike the environmental changes induced for Tyr-16 beta by calcium in the absence of melittin). The conformational changes induced in S100b by interaction with melittin increased its affinity for calcium and offset the inhibition of calcium binding otherwise observed in the presence of potassium ions. This corroborated the previous finding that S100b affinity for calcium greatly depends on the protein conformation. The phenomena described above are similar to the interactions of melittin with calmodulin and thus suggest that S100b and calmodulin have a common structural domain not only that binds melittin but also that may interact with common target proteins.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
Conformational changes in the beta-subunit of the bovine brain Ca2+-binding protein S100b (S100-beta) accompanying Ca2+ binding were investigated by analysis of the spectroscopic properties of the single tyrosine residue (Tyr17 beta) and flow-dialysis binding experiments. S100-beta binds Ca2+ sequentially at two sites to change the conformation of the protein. The first Ca2+ ion binds to site II beta, a typical Ca2+-binding site in the C-terminal region, and it does not significantly perturb the proximal environment of Tyr17 beta. After the first site is occupied, another Ca2+ ion binds to the N-terminal Ca2+-binding site, I beta, and strengthens a hydrogen bond between Tyr17 beta and a neighbouring carboxylate acceptor group, which results in a large increase in the Tyr17 beta fluorescence spectrum half-width and a positive absorption and c.d. signal between 290 and 275 nm. Ca2+ binding to the S100b.Zn2+6 complex, studied by flow-dialysis and fluorescence measurements showed that, although Zn2+ ions increase the affinity of S100b protein for Ca2+, the Ca2+-binding sequence was not changed. Tb3+ (terbium ion) binding studies on the S100b.Zn2+6 complex proved that Tb3+ antagonizes only Ca2+ binding site II beta and confirmed the sequential occupation of Ca2+-binding sites on the S100b.Zn2+6 complex.  相似文献   

4.
Human brain S100b (beta beta) protein was purified using zinc-dependent affinity chromatography on phenyl-Sepharose. The calcium- and zinc-binding properties of the protein were studied by flow dialysis technique and the protein conformation both in the metal-free form and in the presence of Ca2+ or Zn2+ was investigated with ultraviolet spectroscopy, sulfhydryl reactivity and interaction with a hydrophobic fluorescence probe 6-(p-toluidino)naphthalene-2-sulfonic acid (TNS). Flow dialysis measurements of Ca2+ binding to human brain S100b (beta beta) protein revealed six Ca2+-binding sites which we assumed to represent three for each beta monomer, characterized by the macroscopic association constants K1 = 0.44 X 10(5) M-1; K2 = 0.1 X 10(5) M-1 and K3 = 0.08 X 10(5) M-1. In the presence of 120 mM KCl, the affinity of the protein for calcium is drastically reduced. Zinc-binding studies on human S100b protein showed that the protein bound two zinc ions per beta monomer, with macroscopic constants K1 = 4.47 X 10(7) M-1 and K2 = 0.1 X 10(7) M-1. Most of the Zn2+-induced conformational changes occurred after the binding of two zinc ions per mole of S100b protein. These results differ significantly from those for bovine protein and cast doubt on the conservation of the S100 structure during evolution. When calcium binding was studied in the presence of zinc, we noted an increase in the affinity of the protein for calcium, K1 = 4.4 X 10(5) M-1; K2 = 0.57 X 10(5) M-1; K3 = 0.023 X 10(5) M-1. These results indicated that the Ca2+- and Zn2+-binding sites on S100b protein are different and suggest that Zn2+ may regulate Ca2+ binding by increasing the affinity of the protein for calcium.  相似文献   

5.
Flow dialysis measurements of calcium binding to bovine brain S100 alpha alpha, S100a (alpha beta), and S100b (beta beta) proteins in 20 mM Tris-HCl buffer at pH 7.5 and 8.3 revealed that S100 proteins bind specifically 4 Ca2+ eq/mol of protein dimer. The specific calcium-binding sites had, therefore, been assigned to typical amino acid sequences on the alpha and beta subunit. The protein affinity for calcium is much lower in the presence of magnesium and potassium. Potassium strongly antagonizes calcium binding on two calcium-binding sites responsible for most of the Ca2+-induced conformational changes on S100 proteins (probably site II alpha and site II beta). Zinc-binding studies in the absence of divalent cations revealed eight zinc-binding sites/mol of S100b protein dimer that we assumed to correspond to 4 zinc-binding sites/beta subunit. Zinc binding to S100b studied with UV spectroscopy methods showed that the occupation of the four higher affinity sites and the four lower affinity sites on the protein dimer were responsible for different conformational changes in S100b structure. Zinc binding on the higher affinity sites regulates calcium binding to S100b by increasing the protein affinity for calcium and decreasing the antagonistic effect of potassium on calcium binding. Zinc-binding studies on S100a and S100 alpha alpha protein showed that the Trp-containing S100 proteins bind zinc more weakly than S100b protein. Calcium-binding studies on zinc-bound S100a proved that calcium- and zinc-binding sites were distinct although there was no increase in zinc-bound S100a affinity for calcium, as in S100b protein. Finally we provide evidence that discrepancies between previously published results on the optical properties of S100b protein probably result from oxidation of the sulfhydryl groups in the protein.  相似文献   

6.
A rapid separation method for bovine brain S100 alpha alpha, S100a, and S100b protein using fast protein liquid chromatography on a Mono Q column and its application in preparation of a large amount of S100 alpha alpha protein are described. The conformation of S100 alpha alpha in the metal-free forms as well as in the presence of calcium were studied by UV absorption, circular dichroism, intrinsic fluorescence, sulfhydryl reactivity, and interaction with a hydrophobic fluorescent probe. The alpha-subunit appears to have nearly identical conformation in S100 alpha alpha and S100a protein dimers. We also confirmed that only the alpha-subunit exposes hydrophobic domains to solvent in the presence of calcium and that cysteine residues exposed upon Ca2+ binding to S100 proteins correspond to Cys 85 alpha and Cys 84 beta. Incubation of S100a with calcium and KCl proved that calcium binding to the putative calcium-binding sites (site I alpha, I beta) triggers a time- and temperature-dependent conformational change in the protein structure which decreases the antagonistic effect of KCl on calcium binding to sites II alpha and II beta and provokes subunit exchanges between protein dimers and the emergence of S100 alpha alpha and S100b (beta beta) proteins. Dynamic fluorescence measurements showed that incubating calcium at high S100a protein concentrations (greater than 10(-5) M) induces an apparent slow dimer-monomer equilibrium which might result in total subunit dissociation at lower protein concentrations. The effect of acidic pH on subunit dissociation in S100a protein (Morero, R. D., and Weber, G. (1982) Biochim. Biophys. Acta 703, 231-240) arises from conformational changes in the protein structure that are similar to those induced by Ca2+ incubation.  相似文献   

7.
Interactions of trifluoperazine (TFP) with S100 proteins, EF-hand type Ca2+-binding proteins, in the presence of Ca2+ and Zn2+ were studied with induced circular dichroism (CD) and fluorescence spectra. The positive CD bands of TFP were induced at around 265 nm by adding either S100a or S100a0 protein in the presence of Ca2+. No CD band of TFP was, however, induced by adding S100b protein in the presence of Ca2+. Addition of Zn2+ to the TFP/S100 protein solutions did not induce any CD band at all. The fluorescence intensity of 2-p-toluidinylnaphthalene 6-sulfonate (TNS) bound to S100a or S100a0 protein decreased by adding TFP in the presence of Ca2+, while that bound to S100b protein decreased by adding TFP in the presence of Zn2+, indicating that TFP binds to S100a protein and S100a0 protein in a Ca2+-dependent manner and to S100b protein in a Zn2+-dependent manner. From these results together with other experimental findings it was suggested that (1) TFP binds to S100a protein and S100a0 protein in the presence of Ca2+, with half-saturation points of 18 and 3 microM, respectively, (2) TFP binds to S100b protein only in the presence of Zn2+, (3) alpha-subunit of S100 protein binds to TFP specifically in a Ca2+-dependent manner and beta-subunit in a Zn2+-dependent manner.  相似文献   

8.
We purified to homogeneity rat brain S100b protein, which constitutes about 90% of the soluble S100 protein fraction. Purified rat S100b protein comigrates with bovine S100b protein in nondenaturant system electrophoresis but differs in its amino acid composition and in its electrophoretic mobility in urea-sodium dodecyl sulfate-polyacrylamide gel with bovine S100b protein. The properties of the Ca2+ and Zn2+ binding sites on rat S100b protein were investigated by flow dialysis and by fluorometric titration, and the conformation of rat S100b in its metal-free form as well as in the presence of Ca2+ or Zn2+ was studied. The results were compared with those obtained for the bovine S100b protein. In the absence of KCl, rat brain S100b protein is characterized by two high-affinity Ca2+ binding sites with a KD of 2 X 10(-5) M and four lower affinity sites with KD about 10(-4) M. The calcium binding properties of rat S100b protein differ from bovine S100b only by the number of low-affinity calcium binding sites whereas similar Ca2+-induced conformational changes were observed for both proteins. In the presence of 120 mM KCl rat brain S100b protein bound two Zn2+-ions/mol of protein with a KD of 10(-7) M and four other with lower affinity (KD approximately equal to 10(-6) M). The occupancy of the two high-affinity Zn2+ binding sites was responsible for most of the Zn2+-induced conformational changes in the rat S100b protein. No increase in the tyrosine fluorescence quantum yield after Zn2+ binding to rat S100b was observed.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
S100b protein, chemically modified by thioethanol groups (linked via disulfide bonds to two out of four Cys per dimer) was largely similar to reduced native S100b protein in its overall structure and differed only by small modifications extending, however, to the whole protein structure. Studies combining direct Ca2+ binding and associated conformational changes revealed that this chemical modification markedly increased the Ca2(+)-binding affinities (especially in the presence of physiological concentrations of K+ and Mg2+) and introduced a strong positive cooperativity. Different binding models are discussed and it emerges that in both proteins the Ca2(+)-binding sites are not equivalent and probably interact. Like the reduced protein, chemically modified S100b protein binds four Zn2+ ions in two classes of sites (of high and low affinities). Whereas the overall Zn2+ affinity was only slightly decreased, the binding sequence was probably reversed by the introduction of thioethanol groups. Moreover, in the presence of zinc, the Ca2+ affinities were higher and even identical, in both proteins.  相似文献   

10.
The interactions of the S100 protein (S100) with metal cations such as Ca2+, Mg2+, Zn2+ and K+ were studied by the metal n.m.r. spectroscopy. The line widths of 43Ca, 25Mg, 67Zn and 39K n.m.r. markedly increased by adding all S100s. A broad 43Ca n.m.r. band of Ca(2+)-S100a solution was not affected by Zn2+ and K+, while it was greatly decreased by adding Mg2+. The 43Ca n.m.r. spectra of Ca(2+)-S100a0 and -S100b solutions consisted of two slow-exchangeable signals which corresponded to Ca2+ bound to two environmentally different sites of the S100a0. These two 43Ca n.m.r. signals were not affected by Zn2+ and K+. The line width of broad 25Mg n.m.r. band of the Mg(2+)-S100 solution greatly decreased by adding Ca2+, while it did not change by adding Zn2+ and K+. Further, the addition of Ca2+, Mg2+ and K+ did not affect the line width of the 67Zn n.m.r. of the Zn(2+)-S100 solutions. These findings suggest that: (1) Mg2+ binds to all S100s, and at least one of the Mg2+ binding sites of S100 molecule is the same as the Ca2+ binding site; (2) Zn2+ binds to S100s, although the binding site(s) is/are different from Ca(2+)- or Mg(2+)-binding site(s), and the environment of Zn2+ nuclei will not change even though Ca2+ binds to S100s.  相似文献   

11.
19F-n.m.r. spectra were measured to investigate the effects of Ca2+ and Zn2+ on the interaction of trifluoperazine (TFP) with three S100 proteins. It was found that TFP binds to S100a and S100ao proteins irrespective of the presence of Ca2+ and Zn2+, while in the presence of Ca2+ the apparent affinity of TFP to the proteins was greater than that in its absence or in the presence of Zn2+. In contrast, the binding affinity of TRP to S100b protein in the presence and absence of metal ions was lower than to S100a and S100ao proteins. These results suggested that TFP binds to each S100 protein in two ways: one is Ca2(+)- or Zn2(+)-dependent specific manner and another is Ca2(+)- or Zn2(+)-independent non-specific manner.  相似文献   

12.
In addition to binding Ca(2+), the S100 protein S100B binds Zn(2+) with relatively high affinity as confirmed using isothermal titration calorimetry (ITC; K(d) = 94 +/- 17 nM). The Zn(2+)-binding site on Ca(2+)-bound S100B was examined further using NMR spectroscopy and site-directed mutagenesis. Specifically, ITC measurements of S100B mutants (helix 1, H15A and H25A; helix 4, C84A, H85A, and H90A) were found to bind Zn(2+) with lower affinity than wild-type S100B (from 2- to >25-fold). Thus, His-15, His-25, Cys-84, His-85, and perhaps His-90 of S100B are involved in coordinating Zn(2+), which was confirmed by NMR spectroscopy. Previous studies indicate that the binding of Zn(2+) enhances calcium and target protein-binding affinities, which may contribute to its biological function. Thus, chemical shift perturbations observed here for residues in both EF-hand domains of S100B during Zn(2+) titrations could be detecting structural changes in the Ca(2+)-binding domains of S100B that are pertinent to its increase in Ca(2+)-binding affinity in the presence of Zn(2+). Furthermore, Zn(2+) binding causes helix 4 to extend by one full turn when compared to Ca(2+)-bound S100B. This change in secondary structure likely contributes to the increased binding affinity that S100B has for target peptides (i.e., TRTK peptide) in the presence of Zn(2+).  相似文献   

13.
Li T  Hawkes C  Qureshi HY  Kar S  Paudel HK 《Biochemistry》2006,45(10):3134-3145
In the preceding paper, we showed that GSK3beta phosphorylates tau at S(202), T(231), S(396), and S(400) in vivo. Phosphorylation of S(202) occurs without priming. Phosphorylation of T(231), on the other hand, requires priming phosphorylation of S(235). Similarly, priming phosphorylation of S(404) is essential for the sequential phosphorylation of S(400) and S(396) by GSK3beta. The priming kinase that phosphorylates tau at S(235) and S(404) in the brain is not known. In this study, we find that in HEK-293 cells cotransfected with tau, GSK3beta, and Cdk5, Cdk5 phosphorylates tau at S(202), S(235), and S(404). S(235) phosphorylation enhances GSK3beta-catalyzed T(231) phosphorylation. Similarly, Cdk5 by phosphorylating S(404) stimulates phosphorylation of S(400) and S(396) by GSK3beta. These data indicate that Cdk5 primes tau for GSK3beta in intact cells. To evaluate if Cdk5 primes tau for GSK3beta in mammalian brain, we examined localizations of Cdk5, tau, and GSK3beta in rat brain. We also analyzed the interaction of Cdk5 with tau and GSK3beta in brain microtubules. We found that Cdk5, GSK3beta, and tau are virtually colocalized in rat brain cortex. When bovine brain microtubules are analyzed by FPLC gel filtration, Cdk5, GSK3beta, and tau coelute within an approximately 450 kDa complex. From the fractions containing the approximately 450 kDa complex, tau, Cdk5, and GSK3beta co-immunoprecipitate with each other. In HEK-293 cells transfected with tau, Cdk5, and GSK3beta in different combinations, tau binds to Cdk5 in a manner independent of GSK3beta and to GSK3beta in a manner independent of Cdk5. However, Cdk5 and GSK3beta bind to each other only in the presence of tau, suggesting that tau connects Cdk5 and GSK3beta. Our results suggest that in the brain, tau, Cdk5, and GSK3beta are components of an approximately 450 kDa complex. Within the complex, Cdk5 phosphorylates tau at S(235) and primes it for phosphorylation of T(231) by GSK3beta. Similarly, Cdk5 by phosphorylating tau at S(404) primes tau for a sequential phosphorylation of S(400) and S(396) by GSK3beta.  相似文献   

14.
Human S100A2 is an EF-hand calcium-binding S100 protein that is localized mainly in the nucleus and functions as tumor suppressor. In addition to Ca2+ S100A2 binds Zn2+ with a high affinity. Studies have been carried out to investigate whether Zn2+ acts as a regulatory ion for S100A2, as in the case of Ca2+. Using the method of competition with the Zn2+ chelator 4-(2-pyridylazo)-resorcinol, an apparent Kd of 25 nM has been determined for Zn2+ binding to S100A2. The affinity lies close to the range of intracellular free Zn2+ concentrations, suggesting that S100A2 is able to bind Zn2+ in the nucleus. Two Zn2+-binding sites have been identified using site directed mutagenesis and several spectroscopic techniques with Cd2+ and Co2+ as probes. In site 1 Zn2+ is bound by Cys21 and most likely by His 17. The binding of Zn2+ in site 2 induces the formation of a tetramer, whereby the Zn(2+) is coordinated by Cys2 from each subunit. Remarkably, only binding of Zn2+ to site 2 substantially weakens the affinity of S100A2 for Ca2+. Analysis of the individual Ca2+-binding constants revealed that the Ca2+ affinity of one EF-hand is decreased about 3-fold, whereas the other EF-hand exhibits a 300-fold decrease in affinity. These findings imply that S100A2 is regulated by both Zn2+ and Ca2+, and suggest that Zn2+ might deactivate S100A2 by inhibiting response to intracellular Ca2+ signals.  相似文献   

15.
We have recently shown that S-100b protein interacts with the polar surface of cardiolipin vesicles [6]. This interaction produces changes in the secondary structure of S-100b as well as changes in the structural organization of cardiolipin vesicles. We report here on the effects of S-100b on cardiolipin vesicles as investigated by turbidity, terbium-dipicolinate fluorescence and freeze-fracture. Experiments were carried out in the absence and in the presence of Ca2+. In the absence of Ca2+ (0.1 mM EDTA), S-100b favors the aggregation and fusion of vesicles to some extent. Under these conditions, electron microscope analyses reveal the presence of fused vesicles along with particles similar to those observed in protein reconstituted systems or to lipid particles observed during fusional processes. In the presence of Ca2+, S-100b counteracts the Ca2(+)-dependent tendency of vesicles to aggregate and fuse. Under these conditions, bilayer phases along with hexagonal phases can be observed by electron microscopy. The latter effects of S-100b are not due to chelation of Ca2+ because of the relative concentrations of S-100b and Ca2+ under our experimental conditions and since much larger concentrations of EDTA are required to produce the S-100b effects. We propose that the dimeric nature of S-100b plays a major role in these events. In the absence of Ca2+, the S-100b molecules probably cross-link adjacent vesicles, one subunit contacting one vesicle and the other subunit contacting another vesicle through electrostatic bonds. In the presence of Ca2+, due to the large changes occurring in the conformation of the protein (which loses about 52% of its alpha-helical content), S-100b associates strongly with the polar surface of individual vesicles, thus generating some kind of physical barrier to aggregation and fusion of vesicles.  相似文献   

16.
Intra- and Interchain Disulfide Bond Generation in S100b Protein   总被引:1,自引:0,他引:1  
Disulfide-bridged S100b protein formation, aircatalyzed and induced by thiol/disulfide exchange, was studied under various ionic conditions. As native, physiological disulfide-bridged proteins are obtained easily from their reduced counterparts under appropriate redox conditions, this work was performed to determine whether this was the case for disulfide-bridged S100b proteins, reported to have neurite extension activity. In nondenaturating native medium, no disulfide-bridged species could be generated from reduced proteins in any of the ion-induced conformations tested (no ions, Ca2+, Zn2+, or K+) under widely different redox conditions. Only mixed disulfides accumulated, in certain cases. In contrast, intrasubunit monomeric and intersubunit dimeric disulfide-bridged species were readily and efficiently generated under denaturating conditions. A brief characterization of these oxidized species suggested that they differed widely in structure from their reduced counterparts and that they probably did not bind Ca2+. Taken together, these data question the physiological relevance of these disulfide-bridged S100b protein species.  相似文献   

17.
S100B(beta beta) is a dimeric Ca2+-binding protein that interacts with p53, inhibits its phosphorylation by protein kinase C (PKC) and promotes disassembly of the p53 tetramer. Likewise, a 22 residue peptide derived from the C-terminal regulatory domain of p53 has been shown to interact with S100B(beta beta) in a Ca2+-dependent manner and inhibits its phosphorylation by PKC. Hence, structural studies of Ca2+-loaded S100B(beta beta) bound to the p53 peptide were initiated to characterize this interaction. Analysis of nuclear Overhauser effect (NOE) correlations, amide proton exchange rates, 3J(NH-H alpha) coupling constants, and chemical shift index data show that, like apo- and Ca2+-bound S100B(beta beta), S100B remains a dimer in the p53 peptide complex, and each subunit has four helices (helix 1, Glu2-Arg20; helix 2, Lys29-Asn38; helix 3, Gln50-Asp61; helix 4, Phe70-Phe87), four loops (loop 1, Glu21-His25; loop 2, Glu39-Glu49; loop 3, Glu62-Gly66; loop 4, Phe88-Glu91), and two beta-strands (beta-strand 1, Lys26-Lys28; beta-strand 2, Glu67-Asp69), which forms a short antiparallel beta-sheet. However, in the presence of the p53 peptide helix 4 is longer by five residues than in apo- or Ca2+-bound S100B(beta beta). Furthermore, the amide proton exchange rates in helix 3 (K55, V56, E58, T59, L60, D61) are significantly slower than those of Ca2+-bound S100B(beta beta). Together, these observations plus intermolecular NOE correlations between the p53 peptide and S100B(beta beta) support the notion that the p53 peptide binds in a region of S100B(beta beta), which includes residues in helix 2, helix 3, loop 2, and the C-terminal loop, and that binding of the p53 peptide interacts with and induces the extension of helix 4.  相似文献   

18.
The reactivity of sulfhydryl groups of bovine cardiac troponin C   总被引:2,自引:0,他引:2  
Bovine cardiac troponin C (cTnC) contains 2 cysteine residues, Cys-35 located in the nonfunctional Ca2+-binding loop I and Cys-84 in the N-terminal segment of the central helix. We have studied the reactivity of Cys residues in cTnC with 5,5'-dithiobis(2-nitrobenzoic acid) (DTNB) and 7-diethylamino-3-(4'-maleimidylphenyl)-4-methylcoumarin (CPM). The latter compound fluoresces only when reacted with the protein. The reaction with DTNB followed second order kinetics with respect to DTNB, the rate constants being 3.37 s-1 M-1 and 1.82 s-1 M-1 in the presence and absence of Ca2+, respectively. These rates are much slower than the rate of reaction with Cys-98 of skeletal TnC (sTnC) or with the urea-denatured cTnC, indicating that both Cys residues are partly buried within the structure of the protein. The increase in reactivity was induced by binding of Ca2+ to the single low affinity Ca2+ binding site (site II). The fluorescence increase upon reaction of cTnC with CPM in the absence of Ca2+ could be fitted with a single exponential equation indicating that both cysteine residues are equally available to the reagent. The reaction in the presence of Ca2+ was biphasic. Analysis of CNBr fragments of cTnC labeled with CPM under various conditions indicated that in the presence of Ca2+ the reactivity of Cys-84 is increased while that of Cys-35 is slightly decreased. This finding is consistent with the model of Herzberg et al. (Herzberg, O., Moult, J., and James, M. N. G. (1986) J. Biol. Chem. 261, 2638-2644) and the data of Ingraham and Hodges (Ingraham, R. H., and Hodges, R. S. (1988) Biochemistry 27, 5891-5898), suggesting that the Ca2+-induced conformational change in the N-terminal half of TnC involves separation of the helix C from the central helix, thereby increasing the accessibility of Cys-84. The slow overall kinetics, however, indicates that the structure in the vicinity of Cys residues is relatively compact regardless of Ca2+. We interpret the increase in reactivity towards CPM as consistent with a Ca2+-induced exposure of a hydrophobic pocket in the vicinity of Cys-84.  相似文献   

19.
Presynaptic Ca2+ channels are inhibited by metabotropic receptors. A possible mechanism for this inhibition is that G protein betagamma subunits modulate the binding of the Ca2+ channel beta subunit on the Ca2+ channel complex and induce a conformational state from which channel opening is more reluctant. To test this hypothesis, we analyzed the binding of Ca2+ channel beta and G protein beta subunits on the two separate binding sites, i.e. the loopI-II and the C terminus, and on the full-length P/Q-type alpha12.1 subunit by using a modified mammalian two-hybrid system and fluorescence resonance energy transfer (FRET) measurements. Analysis of the interactions on the isolated bindings sites revealed that the Ca2+ channel beta1b subunit induces a strong fluorescent signal when interacting with the loopI-II but not with the C terminus. In contrast, the G protein beta subunit induces FRET signals on both the C terminus and loopI-II. Analysis of the interactions on the full-length channel indicates that Ca2+ channel beta1b and G protein beta subunits bind to the alpha1 subunit at the same time. Coexpression of the G protein increases the FRET signal between alpha1/beta1b FRET pairs but not for alpha1/beta1b FRET pairs where the C terminus was deleted from the alpha1 subunit. The results suggest that the G protein alters the orientation and/or association between the Ca2+ channel beta and alpha12.1 subunits, which involves the C terminus of the alpha1 subunit and may corresponds to a new conformational state of the channel.  相似文献   

20.
Kerkhoff C  Vogl T  Nacken W  Sopalla C  Sorg C 《FEBS letters》1999,460(1):134-138
Analysis of the calcium-induced arachidonic acid (AA) binding to S100A8/A9 revealed that maximal AA binding was achieved at molar ratios of 1 mol S100A8 and 1 mol S100A9 and for values greater than 3 calciums per EF-hand. The AA binding capacity was not induced by the binding of other bivalent cations, such as Zn2+, Cu2+, and Mg2+, to the protein complex. In contrast, the binding of AA was prevented by the addition of either Zn2+ or Cu2+ in the presence of calcium, whereas Mg2+ failed to abrogate the AA binding capacity. The inhibitory effect was not due to blocking the formation of S100A8/A9 as demonstrated by a protein-protein interaction assay. Fluorescence measurements gave evidence that both Zn2+ and Cu2+ induce different conformational changes thereby affecting the calcium-induced formation of the AA binding pocket within the protein complex. Due to the fact that the inhibitory effect of Zn2+ was present at physiological serum concentrations, it is assumed that released S100A8/A9 may carry AA at inflammatory lesions, but not within the blood compartment.  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