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 62 毫秒
1.
A Mn2+-dependent protein phosphatase 2A which is composed of a 34 kDa catalytic C' subunit and a 63 kDa regulatory A' subunit, was purified from human erythrocyte cytosol. C' and A' produced V8- and papain-peptide maps identical to those of the 34 kDa catalytic C and the 63 kDa regulatory A subunits of the Mn2+-independent conventional protein phosphatase in human erythrocyte cytosol, respectively. Reconstitution of C'A and CA' revealed that the metal dependency resided in C' and not in A'. In CA, 0.87 +/- 0.12 mol zinc and 0.35 +/- 0.18 mol iron per mol enzyme were detected by atomic absorption spectrophotometry, but manganese, magnesium and cobalt were not detected. None of these metals was detected in C'A'. Pre-incubation of C' with ZnCl2 and FeCl2, but not FeCl3, synergistically stimulated the Mn2+-independent protein phosphatase activity. The protein phosphatase activity of C was unaffected by the same zinc and/or iron treatment. These results suggest that C is a Zn2+- and Fe2+-metalloenzyme and that C' is the apoenzyme.  相似文献   

2.
A heterodimeric form, CA, of protein-serine/threonine phosphatase (PP) 2A purified from human erythrocytes was dissociated into a 34-kDa catalytic subunit C and 63-kDa inactive subunit A by Sephacryl S-200 gel filtration in the presence of 6 M urea. Reassociation of the C- and A-subunits in the absence of urea suppressed the PP activity of the C subunit toward phosphorylase a, P-H2B histone, and P-H1 histone in the presence or absence of 20 mM MnCl(2) or 50 mM Mg(CH(3)COO)(2), but stimulated the PP activity toward P-H1 histone in the presence of 200 mM NaCl and the Mn(2+)-dependent protein-tyrosine phosphatase (PTP) activity toward P-Tyr-Glu copolymers. The 74-kDa inactive B'(delta) subunit was isolated from a heterotrimeric form, CAB'(delta), of PP2A partially purified from human erythrocytes, by heparin-Sepharose column chromatography. The B'(delta) subunit reassociated with CA and suppressed the PP- and PTP-activities of CA. The B'(delta) subunit did not associate with the isolated C subunit directly, and had no effect on the activities of the C subunit, indicating that the A subunit is essential for the association of the B'(delta) subunit with CA and the resulting suppression of the PP- and PTP-activities.  相似文献   

3.
In vitro protein binding assays identified two distinct calmodulin (CaM) binding sites within the NH(2)-terminal 30-kDa domain of erythrocyte protein 4.1 (4.1R): a Ca(2+)-independent binding site (A(264)KKLWKVCVEHHTFFRL) and a Ca(2+)-dependent binding site (A(181)KKLSMYGVDLHKAKDL). Synthetic peptides corresponding to these sequences bound CaM in vitro; conversely, deletion of these peptides from a 30-kDa construct reduced binding to CaM. Thus, 4.1R is a unique CaM-binding protein in that it has distinct Ca(2+)-dependent and Ca(2+)-independent high affinity CaM binding sites. CaM bound to 4.1R at a stoichiometry of 1:1 both in the presence and absence of Ca(2+), implying that one CaM molecule binds to two distinct sites in the same molecule of 4.1R. Interactions of 4.1R with membrane proteins such as band 3 is regulated by Ca(2+) and CaM. While the intrinsic affinity of the 30-kDa domain for the cytoplasmic tail of erythrocyte membrane band 3 was not altered by elimination of one or both CaM binding sites, the ability of Ca(2+)/CaM to down-regulate 4. 1R-band 3 interaction was abrogated by such deletions. Thus, regulation of protein 4.1 binding to membrane proteins by Ca(2+) and CaM requires binding of CaM to both Ca(2+)-independent and Ca(2+)-dependent sites in protein 4.1.  相似文献   

4.
Recombinant I(1)(PP2A) and I(2)(PP2A) did not affect the activity of the catalytic subunit of protein phosphatase 1 (PP1(C)) with (32)P-labeled myelin basic protein, histone H1, and phosphorylase when assayed in the absence of divalent cations. However, in the presence of Mn(2+), I(1)(PP2A) and I(2)(PP2A) stimulated PP1(C) activity by 15-20-fold with myelin basic protein and histone H1 but not phosphorylase. Half-maximal stimulation occurred at 2 and 4 nM I(1)(PP2A) and I(2)(PP2A), respectively. Moreover, I(1)(PP2A) and I(2)(PP2A) reduced the Mn(2+) requirement by about 30-fold to 10 microM. In contrast, PP1(C) activity was unaffected by I(1)(PP2A) and I(2)(PP2A) in the presence of Co(3+) (0.1 mM), Mg(2+) (2 mM), Ca(2+) (0.5 mM), and Zn(2+) (0.1 mM). Following gel filtration chromatography on Sephacryl S-200 in the presence of Mn(2+), PP1(C) coeluted with I(1)(PP2A) and I(2)(PP2A) in the void volume. However, when I(1)(PP2A) and I(2)(PP2A) or Mn(2+) were omitted, PP1(C) emerged with a V(e)/V(0) of approximately 1.6. The results demonstrate that I(1)(PP2A) and I(2)(PP2A) associate with and modify the substrate specificity of PP1(C) in the presence of physiological concentrations of Mn(2+). A novel role is suggested for I(1)(PP2A) and I(2)(PP2A) in the reciprocal regulation of two major mammalian serine/threonine phosphatases, PP1 and PP2A.  相似文献   

5.
The effects of PK11195, a high-affinity peripheral benzodiazepine receptor (PBR) ligand, on protein phosphorylation in isolated purified rat brain mitochondria were investigated. The isoquinoline carboxamide ligand of PBR, PK11195, but not the benzodiazepine ligand Ro5-4864, in the nanomolar concentration range strongly increased the phosphorylation of 3.5 and 17 kDa polypeptides. The effect of PK11195 was seen in the presence of elevated Ca(2+) levels (3 x 10(-7) to 10(-6) m), but not at very low Ca(2+) levels (10(-8) to 3 x 10(-8) m). This indicates that PBR involves Ca(2+) as a second messenger in the regulation of protein phosphorylation. Staurosporine, an inhibitor of protein kinase activity was able to suppress the PK11195-promoted protein phosphorylation. When the permeability transition pore (PTP) was opened by threshold Ca(2+) load, phosphorylation of the 3.5-kDa polypeptide was diminished, but strong phosphorylation of the 43-kDa protein was revealed. The 43-kDa protein appears to be a PTP-specific phosphoprotein. If PTP was opened, PK11195 did not increase the phosphorylation of the 3.5 and 17-kDa proteins but suppressed the phosphorylation of the PTP-specific 43-kDa phosphoprotein. The ability of PK11195 to increase the protein phosphorylation, which was lost under Ca(2+)-induced PTP opening, was restored again in the presence of calmidazolium, an antagonist of calmodulin and inhibitor of protein phosphatase PP2B. These results show a tight interaction of PBR with the PTP complex in rat brain mitochondria. In conclusion, a novel function of PBR in brain mitochondria has been revealed, and the PBR-mediated protein phosphorylation has to be considered an important element of the PBR-associated signal transducing cascades in mitochondria and cells.  相似文献   

6.
Rat liver sinusoidal endothelial cells (LECs) mediate the removal of hyaluronan (HA) from the circulation via a specific Ca(2+)-independent endocytic receptor. To characterize the receptor biochemically, detergent-soluble extracts were prepared from crude LEC membranes. Using a dot blot assay to quantitate 125I-HA binding activity in CHAPS-solubilized membranes, we detected not only specific Ca(2+)-independent but also specific Ca(2+)-dependent HA-binding activity. Both HA-binding activities behave as integral membrane-associated proteins; they are not released from LEC membranes by treatment at pH 11, and they require detergent for extraction. The Ca(2+)-independent HA receptor was inactivated by treatment at 56 degrees C for 30 min or with 200 mM DTT at 4 degrees C for 30 min, whereas the Ca(2+)-dependent activity actually increased by 75% after treatment at 56 degrees C and only 20% of the Ca(2+)-dependent activity was lost after DTT treatment. A two-cycle membrane extraction protocol using CHAPS partially separated the two HA-binding activities. Eight millimolar KCl and 0.5% CHAPS extracted approximately 50% of the Ca(2+)-independent HA receptor, but only 4-11% of the Ca(2+)-dependent activity. When the KCl and CHAPS concentrations were increased to 2.0 M and 1.5%, respectively, the remaining HA receptor, as well as 89-96% of the Ca(2+)-dependent activity, was then extracted. The Ca(2+)-independent and Ca(2+)-dependent activities could also be further separated using Sephacryl S-400 gel filtration chromatography.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
Calcium-dependent and calcium-independent proteinase activities were detected in extracts of rat ventral prostate and its secretion by use of gelatin-containing SDS-PAGE zymography. Ca(2+)-independent proteinase activities of 22, 26, and 73-79 kDa and Ca(2+)-dependent activities of 58, 63, and 66 kDa were found in the adult gland. The 26- (most intense activity of gland) and 22-kDa activities were present in secretion and were not expressed in the undifferentiated gland of the 10-day-old animal. The Ca(2+)-dependent activities were also present in the secretion, where the 63-kDa form was more prominently expressed than the 58- and 66-kDa bands. The Ca(2+)-dependent and Ca(2+)-independent proteinase activities both responded to a broad range of pH values in the incubation media. The 73-79-kDa Ca(2+)-independent activities were sensitive to benzamidine and the Ca(2+)-dependent activities were inhibited by EDTA and EGTA. Both Ca(2+)-dependent and Ca(2+)-independent proteinase activities responded to androgenic manipulations. Castration was followed by the appearance of a 35-kDa Ca(2+)-independent proteinase (at 2 days) and a 43-kDa Ca(2+)-dependent proteinase (at 4 days). In the Ca(2+)-independent proteinase group, the 73-79-kDa activities were increased somewhat and the 22- and 26-kDa activities decreased after castration. The Ca(2+)-dependent proteinases of 58, 63, and 66 kDa increased in activity with castration, but activity of the 58-kDa form decreased again at 7 days after castration. Treatment of animals upon castration for 4 days with hydrocortisone prevented these changes in proteinase activities whereas treatment with actinomycin D or tranexamic acid did not. Testosterone propionate replacement therapy of rats castrated for 16 days stimulated the activities of the 22- and 26-kDa and 73-79-kDa Ca(2+)-independent and the 58- and 63-kDa Ca(2+)-dependent proteinases with 4 days of therapy. The activities of the 35-kDa Ca(2+)-independent and the 43-kDa Ca(2+)-dependent proteinases were repressed with 8 days of testosterone treatment. Thus, individual proteinases show differential changes in activity during development and in response to androgenic manipulation: this suggests that in addition to proteinases which are secreted, others may be involved in intracellular functions or in mediating tissue organization changes.  相似文献   

8.
9.
The interaction of calmodulin antagonists with a phosphoprotein phosphatase, calcineurin, was investigated using para-nitrophenyl phosphate (pNPP) as a substrate. Calmidazolium, a potent calmodulin antagonist, inhibited the Ni(2+)-stimulated calmodulin-independent phosphatase activity to much the same extent as it did the Ca2+/calmodulin-stimulated activity. Other calmodulin antagonists, such as trifluoperazine, thioridazine, and W-7, also inhibited the Ni(2+)-stimulated phosphatase activity. On the other hand, calmidazolium only weakly and partially inhibited the Mn(2+)-stimulated phosphatase activity and the other calmodulin antagonists examined increased the Mn(2+)-stimulated activity, in the absence of calmodulin. With the addition of an equimolar amount, as to the inhibited holoenzyme, of the purified B subunit of calcineurin, the Ni(2+)-stimulated phosphatase activity recovered from 38 to 63% of the control level in the presence of 5 microM calmidazolium. When the amount of additional B subunit was increased, the phosphatase activity recovered to 94% of the control level, thereby implying that calmidazolium inhibits the Ni(2+)-stimulated phosphatase activity by interacting with the B subunit, in the absence of calmodulin. The Mn(2+)-stimulated phosphatase activity also recovered from the inhibition by calmidazolium, but a much larger amount of the B subunit was necessary for the recovery. These results indicate that the Ni(2+)- and Mn(2+)-stimulated activities of calcineurin are differentially affected by calmodulin antagonists and that the B subunit plays a crucial role in the expression of the Ni(2+)-stimulated phosphatase activity.  相似文献   

10.
Sopina VA 《Tsitologiia》2002,44(11):1120-1128
Tartrate-resistant acid phosphatase (TRAP) of Amoeba proteus (strain B) was represented by 3 of 6 bands (= electromorphs) revealed after disc-electrophoresis in polyacrylamide gels with the use of 2-naphthyl phosphate as a substrate at pH 4.0. The presence of MgCl2, CaCl2 or ZnCl2 (50 mM) in the incubation mixture used for gel staining stimulated activities of all 3 TRAP electromorphs or of two of them (in the case of ZnCl2). When gels were treated with MgCl2, CaCl2 or ZnCl2 (10 and 100 mM, 30 min) before their staining activity of TRAP electromorphs also increased. But unlike 1 M MgCl2 or 1 M CaCl2, 1 M ZnCl2 partly inactivated two of the three TRAP electromorphs. EDTA and EGTA (5 mM), and H2O2 (10 mM) completely inhibited TRAP electromorphs after gel treatment for 10, 20 and 30 min, resp. Of 5 tested ions (Mg2+, Ca2+, Fe2+, Fe3+ and Zn2+), only the latter reactivated the TRAP electromorphs previously inactivated by EDTA or EGTA treatment. In addition, after EDTA inactivation, TRAP electromorphs were reactivated better than after EGTA. The resistance of TRAP electromorphs to okadaic acid and phosphatase inhibitor cocktail 1 used in different concentrations is indicative of the absence of PP1 and PP2A among these electromorphs. Mg2+, Ca2+ and Zn2+ dependence of TRAP activity, and the resistance of its electromorphs to vanadate and phosphatase inhibitor cocktail 2 prevents these electromorphs from being classified as PTP. It is suggested that the active center of A. proteus TRAP contains zinc ion, which is essential for catalytic activity of the enzyme. Thus, TRAP of these amoebae is metallophosphatase showing phosphomonoesterase activity in acidic medium. This metalloenzyme differs from both mammalian tartrate-resistant PAPs and tartrate-resistant metallophosphatase of Rana esculenta.  相似文献   

11.
Insulin-stimulated glucose disposal in skeletal muscle proceeds predominantly through a nonoxidative pathway with glycogen synthase as a rate-limiting enzyme, yet the mechanisms for insulin activation of glycogen synthase are not understood despite years of investigation. Isolation of putative insulin second messengers from beef liver yielded a pseudo-disaccharide consisting of pinitol (3-O-methyl-d-chiro-inositol) beta-1,4 linked to galactosamine chelated with Mn(2+) (called INS2). Here we show that chemically synthesized INS2 has biological activity that significantly enhances insulin reduction of hyperglycemia in streptozotocin diabetic rats. We used computer modeling to dock INS2 onto the known three-dimensional crystal structure of protein phosphatase 2C (PP2C). Modeling and FlexX/CScore energy minimization predicted a unique favorable site on PP2C for INS2 in a surface cleft adjacent to the catalytic center. Binding of INS2 is predicted to involve formation of multiple H-bonds, including one with residue Asp163. Wild-type PP2C activity assayed with a phosphopeptide substrate was potently stimulated in a dose-dependent manner by INS2. In contrast, the D163A mutant of PP2C was not activated by INS2. The D163A mutant and wild-type PP2C in the absence of INS2 had the same Mn(2+)-dependent phosphatase activity with p-nitrophenyl phosphate as a substrate, showing that this mutation did not disrupt the catalytic site. We propose that INS2 allosterically activates PP2C, fulfilling the role of a putative mediator mimetic of insulin signaling to promote protein dephosphorylation and metabolic responses.  相似文献   

12.
A 68-kDa protein that was tyrosine phosphorylated in the presence of Zn2+ and two proteins of 52 and 46 kDa that were tyrosine phosphorylated in the presence of Mg2+ were separated by column chromatography of a sheep platelet high speed supernatant on poly(Glu, Tyr)4:1 copolymer-Sepharose or tyrosine-Sepharose. Phosphorylation of the 68-kDa protein occurred maximally in the presence of Zn2+ while Mg2+ was ineffective. The kinases responsible for the Zn(2+)- and Mg(2+)-dependent tyrosine phosphorylation could also tyrosine phosphorylate poly(Glu, Tyr)4:1, histone, and angiotensin II with the same metal ion specificity. The two tyrosine kinase activities could be also distinguished by their differential response to polyamines and quercetin. Zn2+ stimulation did not appear to be due to the inhibition of a protein tyrosine phosphatase. Sephadex G-100 gel filtration of the fraction showing Zn(2+)-dependent tyrosine phosphorylation of the 68-kDa protein showed that the tyrosine kinase activity corresponded to a molecular mass of 68,000 and it showed a protein band of 68 kDa as detected by silver staining on sodium dodecyl sulfate-polyacrylamide gel.  相似文献   

13.
Ca(2+)-induced inactivation of L-type Ca(2+) is differentially mediated by two C-terminal motifs of the alpha(1C) subunit, L (1572-1587) and K (1599-1651) implicated for calmodulin binding. We found that motif L is composed of a highly selective Ca(2+) sensor and an adjacent Ca(2+)-independent tethering site for calmodulin. The Ca(2+) sensor contributes to higher Ca(2+) sensitivity of the motif L complex with calmodulin. Since only combined mutation of both sites removes Ca(2+)-dependent current decay, the two-site modulation by Ca(2+) and calmodulin may underlie Ca(2+)-induced inactivation of the channel.  相似文献   

14.
Two cellular proteins of 36 and 63 kDa which bind the small T and middle T antigens of polyomavirus recently have been identified as the catalytic and regulatory subunits of the phosphoserine/threonine-specific type 2A protein phosphatase (PP2A). We report here the presence of phosphoseryl phosphatase activity associated with polyomavirus small T and middle T antigens in immunoprecipitates prepared from virus-infected and transformed cells. Phosphatase activity was also found associated with middle T-antigen mutants, some of which had been defined previously to associate with 36- and 63-kDa cellular proteins. Middle T-antigen-associated phosphatase activity was sensitive to okadaic acid and microcystin-LR, inhibitors of PP2A, and insensitive to inhibitor 1 or 2, orthovanadate, or EDTA. Using antiserum specific for the catalytic subunit of PP2A, we found that unlike the majority of PP2A, middle T-antigen-bound PP2A was membrane associated. However, no gross change in the amount, activity, or localization of PP2A could be attributed to middle T-antigen expression in transformed cells. Anti-PP2A antibodies coprecipitated a 63-kDa protein from normal cells and in addition coprecipitated middle T antigen, 60- and 61-kDa proteins (identified as src family members), and an 81-kDa protein from middle T-antigen-transformed cells. Furthermore, we detected protein kinase activity in PP2A immunoprecipitates and protein phosphatase activity in src immune complexes from extracts of middle T-antigen-transformed, but not normal, cells. These results reinforce the notion that at least a portion of middle T antigen bridges a protein kinase with a protein phosphatase.  相似文献   

15.
Two protein tyrosine phosphatase genes, PTP1 and PTP2, are known in Saccharomyces cerevisiae. However, the functions of these tyrosine phosphatases are unknown, because mutations in either or both phosphatase genes have no clear phenotypic effects. In this report, we demonstrate that although ptp2 has no obvious phenotype by itself, it has a profound effect on cell growth when combined with mutations in a novel protein phosphatase gene. Using a colony color sectoring assay, we isolated 25 mutants in which the expression of PTP1 or PTP2 is required for growth. Complementation tests of the mutants showed that they have a mutation in one of three genes. Cloning and sequence determination of one of these gene, PTC1, indicated that it encodes a homolog of the mammalian protein serine/threonine phosphatase 2C (PP2C). The amino acid sequence of the PTC1 product is approximately 35% identical to PP2C. Disruption of PTC1 indicated that the PTC1 function is nonessential. In contrast, ptc1 ptp2 double mutants showed a marked growth defect. To examine whether PTC1 encodes an active protein phosphatase, a glutathione S-transferase (GST)-PTC1 fusion gene was constructed and expressed in Escherichia coli. Purified GST-PTC1 fusion protein hydrolyzed a serine phosphorylated substrate in the presence of the divalent cation Mg2+ or Mn2+. GST-PTC1 also had weak (approximately 0.5% of its serine phosphatase activity) protein tyrosine phosphatase activity.  相似文献   

16.
A [phosphotyrosine]protein phosphatase (PTPPase) was purified almost to homogeneity from rat brain, with [32P]p130gag-fps, an oncogene product of Fujinami sarcoma virus, as substrate. The characteristics of the purified preparation of PTPPase were as follows: the enzyme was a monomer with a molecular mass of 23 kDa; its optimum pH was 5.0-5.5; its activity was not dependent on bivalent cations; its activity was strongly inhibited by sodium vanadate, but was not inhibited by ZnCl2, L(+)-tartrate or NaF; it catalysed the dephosphorylation of [32P]p130gag-fps, [[32P]Tyr]casein, p-nitrophenyl phosphate and L-phosphotyrosine, but did not hydrolyse [[32P]Ser]tubulin, L-phosphoserine, DL-phosphothreonine, 5'-AMP, 2'-AMP or beta-glycerophosphate significantly. During the purification, most of the PTPPase activity was recovered in distinct fractions from those of conventional low-molecular-mass acid phosphatase (APase), which was reported to be a major PTPPase [Chernoff & Li (1985) Arch. Biochem. Biophys. 240, 135-145], from DE-52 DEAE-cellulose column chromatography, and those two enzymes could be completely separated by Sephadex G-75 column chromatography. APase also showed PTPPase activity with [32P]p130gag-fps, but the specific activity was lower than that of PTPPase with molecular mass of 23 kDa, and it was not sensitive to sodium vanadate. These findings suggested that PTPPase (23 kDa) was the major and specific PTPPase in the cell.  相似文献   

17.
18.
In this work, we describe the ability of living cells of Entamoeba histolytica to hydrolyze extracellular ATP. In these intact parasites, whose viability was determined by motility and by the eosin method, ATP hydrolysis was low in the absence of any divalent metal (78 nmol P(i)/h/10(5) cells). Interestingly, in the presence of 5 mM MgCl(2) an ecto-ATPase activity of 300 nmol P(i)/h/10(5) cells was observed. The addition of MgCl(2) to the extracellular medium increased the ecto-ATPase activity in a dose-dependent manner. At 5 mM ATP, half-maximal stimulation of ATP hydrolysis was obtained with 1.23 mM MgCl(2). Both activities were linear with cell density and with time for at least 1 h. The ecto-ATPase activity was also stimulated by MnCl(2) and CaCl(2) but not by SrCl(2), ZnCl(2), or FeCl(3). In fact, FeCl(3) inhibited both Mg(2+)-dependent and Mg(2+)-independent ecto-ATPase activities. The Mg(2+)-independent ATPase activity was unaffected by pH in the range between 6.4 and 8. 4, in which the cells were viable. However, the Mg(2+)-dependent ATPase activity was enhanced concomitantly with the increase in pH. In order to discard the possibility that the ATP hydrolysis observed was due to phosphatase or 5'-nucleotidase activities, several inhibitors for these enzymes were tested. Sodium orthovanadate, sodium fluoride, levamizole, and ammonium molybdate had no effect on the ATPase activities. In the absence of Mg(2+) (basal activity), the apparent K(m) for ATP(4-) was 0.053 +/- 0.008 mM, whereas at saturating MgCl(2) concentrations, the corresponding apparent K(m) for Mg-ATP(2-) for Mg(2+)-dependent ecto-ATPase activity (difference between total and basal ecto-ATPase activity) was 0.503 mM +/- 0.062. Both ecto-ATPase activities were highly specific for ATP and were also able to hydrolyze ADP less efficiently. To identify the observed hydrolytic activities as those of an ecto-ATPase, we used suramin, a competitive antagonist of P(2) purinoreceptors and an inhibitor of some ecto-ATPases, as well as the impermeant agent 4'-4'-diisothiocyanostylbenzene-2'-2'-disulfonic acid. These two reagents inhibited the Mg(2+)-independent and the Mg(2+)-dependent ATPase activities to different extents, and the inhibition by both agents was prevented by ATP. A comparison among the ecto-ATPase activities of three amoeba species showed that the noninvasive E. histolytica and the free-living E. moshkovskii were less efficient than the pathogenic E. histolytica in hydrolyzing ATP. As E. histolytica is known to have a galactose-specific lectin on its surface, which is related to the pathogenesis of amebiasis, galactose was tested for an effect on ecto-ATPase activities. It stimulated the Mg(2+)-dependent ecto-ATPase but not the Mg(2+)-independent ATPase activity.  相似文献   

19.
The functional confirmation of availability of Ca2+ transport initially-active systems in the embryo cells of loach Misgurnus fossilis L. has been obtained. Using thapsigargin, the specific inhibitor of endoplasmic reticulum of Ca2+, Mg(2+)-ATPase, this enzyme activity was divided into thapsigargin-sensitive (actually endoplasmic reticulum Ca2+, Mg(2+)-ATPase) and thapsigargin-insensitive (plasma membrane Ca2+, Mg(2+)-ATPase) constituents. The Ca(2+)-independent Mg(2+)-dependent ATPase activity makes above 39.7% of the common Ca2+, Mg(2+)-ATPase activity of embryo loach. The periodic changes of Ca2+, Mg(2+)-ATPase activity (except for the changes of plasma membrane Ca2+, Mg(2+)-ATPase activity) were found out, which coincide with periodic [Ca2+]i oscillations during the synchronous divisions of loach blastomers embryos.  相似文献   

20.
In this study, we describe the ability of intact fat body of an insect, Rhodnius prolixus, to hydrolyze extracellular ATP. In these fat bodies, the ATP hydrolysis was low in the absence of any divalent metal, and was stimulated by MgCl(2). Both activities (in the absence or presence of MgCl(2)) were linear with time for at least 30 min. In order to confirm the observed nucleotidase activities as ecto-nucleotidases, we used an impermeant inhibitor, DIDS (4, 4'-diisothiocyanostylbene 2'-2'-disulfonic acid). This reagent inhibited both nucleotidase activities and its inhibitory effect was suppressed by ATP. Both ecto-nucleotidase activities were insensitive to inhibitors of other ATPase and phosphatase activities, such as oligomycin, sodium azide, bafilomycin, ouabain, vanadate, molybdate, sodium fluoride, levamizole, tartrate, p-NPP, sodium phosphate, and suramin. Concanavalin A, activator of some ecto-ATPases, was able to stimulate the Mg(2+)-independent nucleotidase activity, but not the Mg(2+)-dependent one. The Mg(2+)-independent nucleotidase activity was enhanced with increases in the pH in the range between 6.4-8.0, but the Mg(2+)-dependent nucleotidase activity was not affected. Besides MgCl(2) , the ecto-ATPase activity was also stimulated by CaCl(2),() MnCl(2), and SrCl(2), but not by ZnCl(2). ATP, ADP, and AMP were the best substrates for the Mg(2+)-dependent ecto-nucleotidase activity, and CTP, GTP, and UTP produced very low reaction rates. However, the Mg(2+)-independent nucleotidase activity recognized all these nucleotides producing similar reaction rates, but GTP was a less efficient substrate. The possible role of the two ecto-nucleotidase activities present on the cell surface of fat body of Rhodnius prolixus, which are distinguished by their substrate specificity and their response to Mg(2+), is discussed.  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