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 9 毫秒
1.
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   

2.
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   

3.
The single tyrosine residue in S-100b protein was nitrated by treatment with tetranitromethane in 0.1 M-Tris/HCl buffer, pH 8.0, containing 2 mM-EDTA. The nitrated protein did not differ significantly in secondary structure from its native unmodified counterpart, as revealed by far-u.v. c.d. measurements. The effect of Ca2+ on the modified protein was different from that on the native protein, e.g. addition of Ca2+ resulted in a loss of helical content from 55 to 47% with the native protein whereas Ca2+ had no significant effect on the gross conformation of the nitrated derivative. Near-u.v. c.d. studies also indicated a very minimal effect on the tyrosine residue and this was also reflected in the u.v.-absorption difference spectrum. Polyacrylamide-gel electrophoresis in the absence of SDS showed the nitrated S-100b to move faster in the presence of EDTA compared with the calcium-bound state, suggesting that the modified protein does bind Ca2+ although it does not undergo a major conformational change in response to Ca2+ addition. In contradistinction, Zn2+ binding was not influenced by nitration, as demonstrated by aromatic c.d. and u.v.-difference spectroscopy. It is clear from this study that the single tyrosine residue in S-100b is critical to sense the Ca2+-induced conformational changes in the protein.  相似文献   

4.
cis-Fatty acids such as oleic acid or linoleic acid have been previously shown to induce full activation of protein kinase C in the absence of Ca2+ and phospholipids (Murakami, K., and Routtenberg, A. (1985) FEBS Lett. 192, 189-193; Murakami, K., Chan, S.Y., and Routtenberg, A. (1986) J. Biol. Chem. 261, 15424-15429). In this study, we have investigated the effects of various metal ions on protein kinase C activity without the interference of Ca2+ since cis-fatty acid requires no Ca2+ for protein kinase C activation. Here we report a specific interaction of Zn2+ with protein kinase C in either a positive or negative cooperative fashion in concert with Ca2+. At low concentrations (approximately 5 microM) of Ca2+, Zn2+ enhances protein kinase C activity induced by both oleic acid and phosphatidylserine/diolein. In contrast, Zn2+ inhibits the activity at higher concentrations (over 50 microM) of Ca2+. In the absence of Ca2+, Zn2+ shows no effect on protein kinase C activity. Our results suggest that Zn2+ does not recognize or interact with protein kinase C in the absence of Ca2+, that protein kinase C possesses high and low affinity Ca2+-binding sites, and that at least one Zn2+-binding site exists which is distinct from Ca2+-binding sites.  相似文献   

5.
6.
The 19F n.m.r. spectrum of 3-fluoro-3-deoxy-D-glucose (3FG) in a red-cell suspension was observed to contain separate resonances from the intra- and extra-cellular populations of both the alpha- and beta-anomers. This phenomenon was used with an n.m.r. spin-transfer procedure to measure the rate of exchange of the anomers across the human red-cell membrane under equilibrium-exchange conditions at 37 degrees C. The beta-anomer crossed the membrane significantly more quickly than the alpha-anomer. At a total 3FG concentration of 9.3 mM; the first-order rate constants for the efflux of the alpha- and beta-anomers were 0.41 +/- 0.15 and 0.88 +/- 0.20 s-1 respectively. The measurable 3FG exchange was inhibited by 75 and 100% respectively by the glucose-transport inhibitors cytochalasin B and phloretin. Glucose inhibited the exchange of 3FG, and the results were consistent with glucose and 3FG binding to the hexose-transport protein with similar affinity.  相似文献   

7.
S100 proteins (16 members) show a very divergent pattern of cell- and tissue-specific expression, of subcel-lular localizations and relocations, of post-translational modifications, and of affinities for Ca 2+ , Zn 2+ , and Cu 2+ , consistent with their pleiotropic intra- and extracellular functions. Up to 40 target proteins are reported to interact with S100 proteins and for S100A1 alone 15 target proteins are presently known. Therefore it is not surprising that many functional roles have been proposed and that several human disorders such as cancer, neurodegenerative diseases, cardiomyopathies, inflammations, diabetes, and allergies are associated with an altered expression of S100 proteins. It is not unlikely that their biological activity in some cases is regulated by Zn 2+ and Cu 2+ , rather than by Ca 2+ Despite the numerous putative functions of S100 proteins, their three-dimensional structures of, e.g., S100B, S100A6, and S100A7 are surprisingly similar. They contain a compact dimerization domain whose conformation is rather insensitive to Ca 2+ binding and two lateral a-helices III and III, which project outward of each subunit when Ca 2+ is bound. Target docking depends on the two hydrophobic patches in front of the paired EF-hand generated by the binding of Ca 2+. The selec-tivity in target binding is assured by the central linker between the two EF-hands and the C-terminal tail. It appears that the S100-binding domain in some target proteins contains a basic amphiphilic a-helix and that the mode of interaction and activation bears structural similarity to that of calmodulin.© Kluwer Academic Publishers  相似文献   

8.
S100a is a heterodimeric, acidic calcium-binding protein that interacts with calmodulin antagonists in a Ca2+-dependent manner. In order to study the behavior of the hydrophobic domain on S100a when bound to Ca2+, its interaction with trifluoperazine (TFP) was investigated using16F nuclear magnetic resonance (NMR) and circular dichroism (CD) spectroscopy. The dissociation constant (K d) values of TFP, as estimated from the chemical shifts of19F NMR, were 191 and 29 μm in the absence and presence of Ca2+, respectively, and were similar to those previously reported for S100b. However, the TFP linewidth in the presence of Ca2+-bound S100a was 65 Hz greater than in the presence of Ca2+-bound S100b. This suggests a slower TFP exchange rate for S100a than for S100b. Thus, the TFP linewidths observed for each isoform may reflect differences in structural and modulatory properties of the Ca2+-dependent hydrophobic domains on S100a and S100b. Additionally, the presence of magnesium had no effect on the observed Ca2+-induced TFP spectral changes in S100a solutions. Circular dichroism studies indicate that Ca2+ induces a small transition from α-helix to random coil in S100a; in contrast, the opposite transition is reported for calmodulin (Hennesseyet al., 1987). However, TFP did not significantly alter the secondary structure of Ca2+-bound S100a; this observation is similar to the effect of TFP on Ca2+-bound calmodulin and troponin C (Shimizu and Hatano, 1984; Gariépy and Hodges, 1983). It is, therefore, proposed that TFP binds to a hydrophobic domain on S100a in a fashion similar to other calcium-modulated proteins.  相似文献   

9.
The crystal structure of human psoriasin (S100A7) in the native, calcium-bound form has been determined from two crystal forms of the protein crystallized with and without divalent zinc. The overall structures of the dimeric protein closely resemble the previously determined holmium-substituted structure. The structures also reveal a zinc-binding site of the protein, which is formed by three histidines and an aspartate residue. Together, these residues coordinate the zinc ion in a way similar to the pattern seen in certain metalloproteases and in particular the collagenase family of proteins. Sequence comparison suggests that this zinc site is present in a number of the remaining members of the S100 family. The structure of S100A7 crystallized in the absence of zinc further shows that loss of zinc results in a reorganization of the adjacent empty and distorted EF-hand loop, causing it to resemble a calcium-loaded EF-hand.  相似文献   

10.
This paper presents the first unambiguous demonstration that a unique protein isolated from the hemolysate of human erythrocytes is responsible for increasing both the apparent Ca2+ ion affinity and maximum rate of ATP hydrolysis of the membrane-bound Ca2+Mg2+ ATPase. Unlike previous reports where an unpurified extract from red blood cells was used to activate the ATPase, our results clearly demonstrate that a single protein species, whether initially associated with or added back to the membrane is responsible for the observed changes in ATPase activity.  相似文献   

11.
The calcyclin-like protein from Ehrlich ascites tumour cells is a 10.5 kDa heat stable protein, which binds two Ca2+ ions each with different affinity. Upon Ca2+ binding, the protein changes its conformation exposing hydrophobic regions. In this conformation it is able to interact with fluphenazine and with a 36 kDa protein immunologically similar to mammalian calpactin. Calcyclin-like protein binds Zn2+ and forms dimers like other members of the S-100 protein family. The calcyclin-like protein is present in several mouse tissues such as stomach, skeletal muscle, heart, spleen, lung and kidney, but seems to be absent from brain, intestine and liver as well as from some tumorigenic cells lines.  相似文献   

12.
The S100 genes encode a conserved group of 21 vertebrate‐specific EF‐hand calcium‐binding proteins. Since their discovery in 1965, S100 proteins have remained enigmatic in terms of their cellular functions. In this review, we summarize the calcium‐ and zinc‐binding properties of the dimeric S100B and S100A1 proteins and highlight data that shed new light on the extracellular and intracellular regulation and functions of S100B. We point out that S100B and S100A1 homodimers are not functionally interchangeable and that in a S100A1/S100B heterodimer, S100A1 acts as a negative regulator for the ability of S100B to bind Zn2+. The Ca2+ and Zn2+‐dependent interactions of S100B with a wide array of proteins form the basis of its activities and have led to the derivation of some initial rules for S100B recognition of protein targets. However, recent findings have strongly suggested that these rules need to be revisited. Here, we describe a new consensus S100B binding motif present in intracellular and extracellular vertebrate‐specific proteins and propose a new model for stable interactions of S100B dimers with full‐length target proteins. A chaperone‐associated function for intracellular S100B in adaptive cellular stress responses is also discussed. This review may help guide future studies on the functions of S100 proteins in general.  相似文献   

13.
Resonance Energy Transfer between N-(7-nitro-2,1,3 benzoxadiazol -4 yl) phosphatidyl ethanolamine and N-Lissamine-Rhodamine B sulfonyl) phosphatidyl ethanolamine embedded in two different populations of small unilamellar vesicles made of phosphatidyl serine has been used to study the fusion process induced by Zn2+ and Ca2+. Lipid intermixing demonstrating fusion of liposome membranes can already be observed at 125 and 250 mumol/l of Zn2+. After short time pre-incubations with micromolar concentrations of Zn2+ as low as 150 mumol/l, Ca2+ induces an instantaneous increase of vesicle fusion. The lipid intermixing induced by micromolar concentrations of Ca2+ (250-500 mumol/l) could be increased up to 4 times when pre-incubated with 150 or 200 mumol/l of Zn2+. The effect of 1 mM of Ca2+ alone on lipid intermixing can be mimicked by 150 mumol/l of Zn2+ followed by 500 mumol/l of Ca2+. Our data demonstrate that Zn2+ and Ca2+ act synergistically to affect cation-induced membrane fusion. We suggest that Zn2+ specifically alters the physical state of phospholipid membranes making them more prone to calcium-triggered fusion.  相似文献   

14.
Myelin basic proteins (MBP) interacts with F-actin resulting in the precipitation of a complex of both proteins. Electron microscope observations of this complex reveal the presence of ordered bundles of F-actin filaments similar to those obtained from F-actin and troponin I. In addition to the bundles, there also appear short fragments of F-actin filaments. In the presence of Ca2+ calmodulin causes a release of MBP from its complex with F-actin, accompanied by dissociation of F-actin bundles into separate filaments. Parallel to the binding of MBP to F-actin the ATPase activity of actomyosin is progressively reduced. This inhibition is reversed by calmodulin but only in the presence of Ca2+. Studies of the binding of S-1 to F-actin and to the F-actin-MBP complex indicate that the interaction sites for MBP and S-1 on the actin molecule are different.  相似文献   

15.
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   

16.
Ca2+-binding of S-100 protein was studied using a Ca2+ electrode at pH 6.80. In the presence of 0.1 M KCl and 10 mM MgCl2 (ionic strength 0.13), Ca2+-binding to S-100 protein occurred in three steps with positive cooperativity. The numbers of bound Ca2+ ions in the three steps were 2, 2, and 4. The Ca2+-binding constants were 6.9 x 10(3) M-1, 2.9 x 10(3) M-1, and 3.7 x 10(2) M-1, respectively. The Ca2+-binding constants of the first and second steps obtained in the presence of 33.3 mM MgCl2 or 0.1 M KCl (ionic strength 0.10) were 1.4 times larger than those described above. This suggests that Mg2+ does not inhibit Ca2+-binding of S-100 protein. The increase of KCl concentration from 0.1 to 0.2 M caused a decrease of the Ca2+-binding constants to ca. 50%.  相似文献   

17.
19F-n.m.r. spectroscopy was used to study the binding of 3',5'-difluoromethotrexate to dihydrofolate reductase (tetrahydrofolate dehydrogenase) from Lactobacillus casei. The benzoyl ring of the bound difluoromethotrexate was found to 'flip' about its symmetry axis, and the rate (7.3 X 10(3) s-1 at 298 K) and activation parameters for this process were determined by lineshape analysis of the 19F-n.m.r. spectrum at a series of temperatures in the range 273-308 K. The contributions to the barrier for this process are discussed. Addition of NADP+ or NADPH to form the enzyme-difluoromethotrexate-coenzyme ternary complex led to an increase in the rate of benzoyl ring flipping by a factor of 2.6-2.8-fold, and to substantial changes in the 19F-n.m.r. chemical shifts. The possible nature of the coenzyme-induced conformational changes responsible for these effects is discussed.  相似文献   

18.
The coupling of aliphatic amines to agarose by the cyanogen bromide reaction yields isourea linkages which are positively charged at pH 7. The presence of these cationic sites in affinity gels causes significant non-specific adsorption of proteins. Serum albumin was found to bind to a number of derivatized gels which possessed these charged groups. The use of adipic dihydrazide as the leash moiety yielded affinity gels which were noncharged at pH 7. Serum albumin failed to adsorb to these gels. Beta-galactosidase from Escherichia coli was found to be sensitive to both ionic and hydrophobic groups in an affinity gel. A sample of active-site inhibited enzyme was found to bind to an affinity gel which contained both the cationic isourea and a phenyl structure in the leash. Thus it was concluded that the affinity purification of this enzyme has yet to be demonstrated. These studies dictate against the use of salt and pH gradients to desorb enzymes from affinity sorbents.  相似文献   

19.
20.
11B n.m.r. spectroscopy studies of solutions of sugars in the presence of borax have shown the existence of boron-containing complexes. Chemical-shift values indicate the size of the ring in which boron is involved. Use of this method and p.m.r. spectroscopy showed that D-glucose forms a 1,2-furanoid and 1,2-pyranoid complex in the presence of borax and benzeneboronic acid.  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