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 31 毫秒
1.
Rabbit iris smooth muscle was prelabelled with myo-[3H]inositol for 90 min and the effect of carbachol on the accumulation of inositol phosphates from phosphatidylinositol 4,5-bisphosphate [PtdIns(4,5)P2], phosphatidylinositol 4-phosphate (PtdIns4P) and phosphatidylinositol (PtdIns) was monitored with anion-exchange chromatography. Carbachol stimulated the accumulation of inositol phosphates and this was blocked by atropine, a muscarinic antagonist, and it was unaffected by 2-deoxyglucose. The data presented demonstrate that, in the iris, carbachol (50 microM) stimulates the rapid breakdown of PtdIns(4,5)P2 into [3H]inositol trisphosphate (InsP3) and diacylglycerol, measured as phosphatidate, and that the accumulation of InsP3 precedes that of [3H]inositol bisphosphate (InsP2) and [3H]inositol phosphate (InsP). This conclusion is based on the following findings. Time course experiments with myo-[3H]inositol revealed that carbachol increased the accumulation of InsP3 by 12% in 15s and by 23% in 30s; in contrast, a significant increase in InsP release was not observed until about 2 min. Time-course experiments with 32P revealed a 10% loss of radioactivity from PtdIns(4,5)P2 and a corresponding 10% increase in phosphatidate labelling by carbachol in 15s; in contrast a significant increase in PtdIns labelling occurred in 5 min. Dose-response studies revealed that 5 microM-carbachol significantly increased (16%) the accumulation of InsP3 whereas a significant increase in accumulation of InsP2 and InsP was observed only at agonist concentrations greater than 10 microM. Studies on the involvement of Ca2+ in the agonist-stimulated breakdown of PtdIns(4,5)P2 in the iris revealed the following. Marked stimulation (58-78%) of inositol phosphates accumulation by carbachol in 10 min was observed in the absence of extracellular Ca2+. Like the stimulatory effect of noradrenaline, the ionophore A23187-stimulated accumulation of InsP3 was inhibited by prazosin, an alpha 1-adrenergic blocker, thus suggesting that the ionophore stimulation of PtdIns(4,5)P2 breakdown we reported previously [Akhtar & Abdel-Latif (1978) J. Pharmacol. Exp. Ther. 204, 655-688; Akhtar & Abdel-Latif (1980) Biochem. J. 192, 783-791] was secondary to the release of noradrenaline by the ionophore. The carbachol-stimulated accumulation of inositol phosphates was inhibited by EGTA (0.25 mM) and this inhibition was reversed by excess Ca2+ (1.5 mM), suggesting that EGTA treatment of the tissue chelates extracellular Ca2+ required for polyphosphoinositide phosphodiesterase activity. K+ depolarization, which causes influx of extracellular Ca2+ in smooth muscle, did not change the level of InsP3.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

2.
Phosphatidylinositol (PtdIns), phosphatidylinositol 4-phosphate (PtdIns4P) and phosphatidylinositol 4,5-bisphosphate [PtdIns(4,5)P2] of turkey erythrocytes were labelled by using either [32P]Pi or [3H]inositol. Although there was little basal release of inositol phosphates from membranes purified from labelled cells, in the presence of guanosine 5'-[gamma-thio]triphosphate (GTP[S]) the rate of accumulation of inositol bis-, tris- and tetrakis-phosphate (InsP2, InsP3 and InsP4) was increased 20-50-fold. The enhanced rate of accumulation of 3H-labelled inositol phosphates was linear for up to 20 min; owing to decreases in 32P specific radioactivity of phosphoinositides during incubation of membranes with unlabelled ATP, the accumulation of 32P-labelled inositol phosphates was linear for only 5 min. In the absence of ATP and a nucleotide-regenerating system, no InsP4 was formed, and the overall inositol phosphate response to GTP[S] was decreased. Analyses of phosphoinositides during incubation with ATP indicated that interconversions of PtdIns to PtdIns4P and PtdIns4P to PtdIns(4,5)P2 occurred to maintain PtdIns(4,5)P2 concentrations; GTP[S]-induced inositol phosphate formation was accompanied by a corresponding decrease in 32P- and 3H-labelled PtdIns, PtdIns4P and PtdIns(4,5)P2. In the absence of ATP, only GTP[S]-induced decreases in PtdIns(4,5)P2 occurred. Since inositol monophosphate was not formed under any condition, PtdIns is not a substrate for the phospholipase C. The production of InsP2 was decreased markedly, but not blocked, under conditions where Ins(1,4,5)P3 5-phosphomonoesterase activity in the preparation was inhibited. Thus the predominant substrate of the GTP[S]-activated phospholipase C of turkey erythrocyte membranes is PtdIns(4,5)P2. Ins(1,4,5)P3 was the major product of this reaction; only a small amount of Ins(1:2-cyclic, 4,5)P3 was released. The effects of ATP on inositol phosphate formation apparently involve the contributions of two phenomena. First, the P2-receptor agonist 2-methylthioadenosine triphosphate (2MeSATP) greatly increased inositol phosphate formation and decreased [3H]PtdIns4P and [3H]PtdIns(4,5)P2 in the presence of a low (0.1 microM) concentration of GTP[S]. ATP over the concentration range 0-100 microM produced effects in the presence of 0.1 microM-GTP[S] essentially identical with those observed with 2MeSATP, suggesting that the effects of low concentrations of ATP are also explained by a stimulation of P2-receptors. Higher concentrations of ATP also increase inositol phosphate formation, apparently by supporting the synthesis of substrate phospholipids.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

3.
When hepatocytes were incubated with [32P]Pi, the kinetics for the labelling of the monoester phosphate groups of phosphatidylinositol 4-phosphate and phosphatidylinositol 4,5-bisphosphate were similar to each other and slightly slower than that for the labelling of the gamma-phosphate of ATP. Analysis of the water-soluble 3H-labelled materials derived from [3H]inositol-labelled hepatocytes revealed that, in addition to inositol and its mono-, bis- and tris-phosphates (Ins, InsP, InsP2 and InsP3), these cells contained two unidentified radioactive compounds which co-eluted with InsP on anion-exchange chromatography. When [3H]inositol-labelled hepatocytes were stimulated with 0.23 microM-vasopressin in the presence of 10 mM-Li+, there was an accumulation of radioactivity in InsP, InsP2 and InsP3 but not in Ins or the two unidentified compounds. Further analysis of these inositol phosphates by h.p.l.c. revealed that vasopressin also stimulates the accumulation of inositol tetrakisphosphate (InsP4) in these cells. Vasopressin-stimulated InsP and InsP2 accumulations were maximal in the presence of 1-10 mM-Li+ but InsP3 accumulation continued to increase up to 50 mM-Li+. Accumulated inositol phosphates were retained within the cell. Li+ from 1 to 50 mM did not influence the extent of vasopressin-stimulated inositol lipid degradation in hepatocytes. In the absence of Li+, radioactivity in vasopressin-stimulated hepatocytes accumulated almost entirely in free inositol. The vasopressin-stimulated accumulation of inositol phosphates in the presence of 10 mM-Li+ was abolished by a V1-vasopressin antagonist. Inositol phosphate accumulation was not influenced by ionophore A23187, dimethyl sulphoxide or indomethacin.  相似文献   

4.
Experiments with washed rabbit platelets demonstrate that stimulation with a low concentration of thrombin (0.1 unit/ml), that causes maximal aggregation and partial release of amine granule contents, also causes increased accumulation of [3H]inositol-labelled inositol trisphosphate (InsP3) in the presence of 20 mM-Li+. This concentration of Li+ was found to inhibit the degradation of inositol phosphates by phosphomonoesterases. This result indicates that phosphatidylinositol 4,5-bisphosphate [PtdIns(4,5)P2] is degraded early after platelet stimulation with thrombin, although in a previous study we had found no decrease in amount. In the absence of Li+, the labelling of inositol bisphosphate (InsP2) increased more rapidly than that of InsP3, consistent with rapid degradation of InsP3 by phosphomonoesterase. After 30s the increase in InsP2 was augmented by Li+. This increase in InsP2 could have been due to increased degradation of phosphatidylinositol 4-phosphate or inhibition of breakdown of InsP2 to InsP with a lesser inhibition of breakdown of InsP3 to InsP2. The effect on InsP3 and InsP2 of stimulation of the platelets with 1.0 unit of thrombin/ml was comparable with the effect of the lower concentration of thrombin. Inositol phosphate (InsP) labelling did not increase in response to 0.1 unit of thrombin/ml, but increased when the platelets were stimulated with 1.0 unit of thrombin/ml. Whether the increase in InsP was due to increased degradation of phosphatidylinositol or a greater rate of breakdown of InsP2 to InsP than InsP to inositol cannot be determined in these experiments. These results indicate that degradation of PtdIns(4,5)P2 is an early event in platelet activation by thrombin and that formation of inositol phosphates and 1,2-diacylglycerol rather than a decrease in PtdIns(4,5)P2 may be the important change.  相似文献   

5.
The effects of the muscarinic agonist carbachol, histamine and bradykinin on incorporation of [3H]inositol into the phosphoinositides and the formation of [3H]InsPs were examined in bovine tracheal smooth-muscle (BTSM) slices labelled with [3H]inositol. These agonists result in substantial and dose-related increases in the incorporation of [3H]inositol into the phospholipids. Carbachol and histamine stimulated the incorporation of [3H]inositol into the phospholipids to the same degree, despite histamine being only 35% as effective as carbachol on [3H]InsP accumulation. Histamine and carbachol, at maximal concentrations, were non-additive with respect to both the stimulated incorporation of [3H]inositol and [3H]InsP formation. For carbachol this effect on incorporation was found to occur to a similar extent in PtdInsP and PtdInsP2 as well as PtdIns. The initial effect of carbachol on [3H]inositol incorporation was rapid (maximal by 10 min); however, with prolonged stimulation large secondary declines in PtdInsP and PtdInsP2 labelling were observed, with depletion of the much larger PtdIns pool only evident in the presence of Li+. Lowering buffer [Ca2+] increased the incorporation of [3H]inositol under basal conditions, but did not attenuate the subsequent agonist-stimulated incorporation effect. The large changes in specific radioactivity of the phosphoinositides, and consequently the [3H]InsP products, after carbachol stimulation resulted in the apparent failure of atropine to reverse the [3H]InsP response completely. Labelling muscle slices with [3H]inositol in the presence of carbachol or labelling for longer periods (greater than 6 h) prevented subsequent carbachol-stimulated effects on incorporation without significantly altering the dose-response relationship for carbachol-stimulated [3H]InsP formation and resulted in steady-state labelling conditions confirmed by the ability of atropine to reverse fully the [3H]InsP response to carbachol. This study demonstrates the profound effects of a number of agonists on [3H]inositol incorporation into the phospho- and polyphosphoinositides in BTSM with important consequent changes in the specific radioactivity of these lipids and the resulting [3H]InsP products. In addition, a selective depletion of PtdInsP and PtdInsP2 over PtdIns has been demonstrated with prolonged muscarinic-receptor stimulation.  相似文献   

6.
Ca2+ efficiently inhibits binding of inositol 1,4,5-trisphosphate (InsP3) to the InsP3 receptor in cerebellar membranes but not to the purified receptor. We have now investigated the mechanism of action by which Ca2+ inhibits InsP3 binding. Our results suggest that Ca2+ does not cause the stable association of a Ca(2+)-binding protein with the receptor. Instead, Ca2+ leads to the production of a soluble, heat-stable, low molecular weight substance from cerebellar membranes that competes with InsP3 for binding. This inhibitory substance probably represents endogenously generated InsP3 as judged by the fact that it co-purifies with InsP3 on anion-exchange chromatography, competes with [3H]InsP3 binding in a pattern similar to unlabeled InsP3, and is in itself capable of releasing 45Ca2+ from permeabilized cells. A potent Ca(2+)-activated phospholipase C activity producing InsP3 was found in cerebellar microsomes that exhibited a Ca2+ dependence identical to the Ca(2+)-dependent inhibition of InsP3 binding. Together these results suggest that the Ca(2+)-dependent inhibition of InsP3 binding to the cerebellar receptor is due to activation of a Ca(2+)-sensitive phospholipase C enriched in cerebellum. Nevertheless, Ca2+ probably also modulates the InsP3 receptor function by a direct interaction with the receptor that does not affect InsP3 binding but regulates InsP3-dependent channel gating.  相似文献   

7.
WRK 1 cells were labelled to equilibrium with 2-myo-[3H]inositol and stimulated with vasopressin. Within 3 s of hormone stimulation there was a marked accumulation of 3H-labelled InsP2 and InsP3 (inositol bis- and tris-phosphate), but not of InsP (inositol monophosphate). There was an associated, and rapid, depletion of 3H-labelled PtdInsP and PtdInsP2 (phosphatidylinositol mono- and bis-phosphates), but not of PtdIns (phosphatidylinositol), in these cells. Some 4% of the radioactivity in the total inositol lipid pool of WRK 1 cells was recovered in InsP2 and InsP3 after 10 s stimulation with the hormone. The selectivity of the vasopressin receptors of WRK 1 cells for a variety of vasopressin agonists and antagonists revealed these to be of the V1a subtype. There was no receptor reserve for vasopressin-stimulated inositol phosphate accumulation in WRK 1 cells. The accumulation of inositol phosphates was enhanced in the presence of Li+ions. Half-maximal accumulation of InsP, InsP2 and InsP3 in vasopressin-stimulated cells was observed with 0.9, 3.0 and 3.6 mM-Li+ respectively. Bradykinin and 5-hydroxytryptamine also provoked inositol phosphate accumulation in WRK 1 cells. The effects of sub-optimal concentrations of bradykinin and vasopressin upon inositol phosphate accumulation were additive, but those of optimal concentrations of the hormones were not.  相似文献   

8.
3-phosphoinositide-dependent protein kinase-1 (PDK1) phosphorylates and activates many kinases belonging to the AGC subfamily. PDK1 possesses a C-terminal pleckstrin homology (PH) domain that interacts with PtdIns(3,4,5)P3/PtdIns(3,4)P2 and with lower affinity to PtdIns(4,5)P2. We describe the crystal structure of the PDK1 PH domain, in the absence and presence of PtdIns(3,4,5)P3 and Ins(1,3,4,5)P4. The structures reveal a 'budded' PH domain fold, possessing an N-terminal extension forming an integral part of the overall fold, and display an unusually spacious ligand-binding site. Mutagenesis and lipid-binding studies were used to define the contribution of residues involved in phosphoinositide binding. Using a novel quantitative binding assay, we found that Ins(1,3,4,5,6)P5 and InsP6, which are present at micromolar levels in the cytosol, interact with full-length PDK1 with nanomolar affinities. Utilising the isolated PDK1 PH domain, which has reduced affinity for Ins(1,3,4,5,6)P5/InsP6, we perform localisation studies that suggest that these inositol phosphates serve to anchor a portion of cellular PDK1 in the cytosol, where it could activate its substrates such as p70 S6-kinase and p90 ribosomal S6 kinase that do not interact with phosphoinositides.  相似文献   

9.
InsP(6) is abundant in cereals and legumes. InsP(6) and lower inositol phosphates, in particular InsP(3), participate in important intracellular processes. In addition, InsP(6) possess significant health benefits, such as anti-cancer effect, kidney stones prevention, lowering serum cholesterol. Because of the insensitivity of existing methods for determination of non-radiolabeled inositol phosphates, little is known about the natural occurrence, much less on the concentrations of InsP(6) and InsP(3) in biological samples. Using gas chromatography-mass detection analysis of HPLC chromatographic fractions, we report a measurement of unlabeled total InsP(3) and InsP(6) (a) as they occur within cells culture, tissues, and plasma, and (b) their changes depending on the presence of exogenous InsP(6). When rats were fed on a purified diet in which InsP(6) was undetectable (AIN-76A) the levels of InsP(6) in brain were 3.35 +/- 0.57 (SE) micromol.kg(-1) and in plasma 0.023 +/- 0.008 (SE) micromol.l(-1). The presence of InsP(6) in diet dramatically influenced its levels in brain and in plasma. When rats were given an InsP(6)-sufficient diet (AIN-76A + 1% InsP(6)), the levels of InsP(6) were about 100-fold higher in brain tissues (36.8 +/- 1.8 (SE)) than in plasma (0.29 +/- 0.02 (SE)); InsP(6) concentrations were 8.5-fold higher than total InsP(3) concentrations in either plasma (0.033 +/- 0.012 (SE)) and brain (4.21 +/- 0.55 (SE)). When animals were given an InsP(6)-poor diet (AIN-76A only), there was a 90% decrease in InsP(6) content in both brain tissue and plasma (p < 0.001); however, there was no change in the level of total InsP(3). In non-stimulated malignant cells (MDA-MB 231 and K562) the InsP(6) contents were 16.2 +/- 9.1 (SE) micromol.kg(-1) for MDA-MB 231 cells and 15.6 +/- 2.7 (SE) for K 562 cells. These values were around 3-fold higher than those of InsP(3) (4.8 +/- 0.5 micromol.kg(-1) and 6.9 +/- 0.1 (SE) for MDA-MB 231 and K562 cells respectively). Treatment of malignant cells with InsP(6) resulted in a 2-fold increase in the intracellular concentrations of total InsP(3) (9.5 +/- 1.3 (SE) and 10.8 +/- 1.0 (SE) micromol.kg(-1) for MDA-MB 231 and K562 cells respectively, p < 0.05), without changes in InsP(6) levels. These results indicate that exogenous InsP(6) directly affects its physiological levels in plasma and brain of normal rats without changes on the total InsP(3) levels. Although a similar fluctuation of InsP(6) concentration was not seen in human malignant cell lines following InsP(6) treatment, an increased intracellular levels of total InsP(3) was clearly observed.  相似文献   

10.
Stimulation of the hydrolysis of phosphatidylinositol 4,5-bisphosphate (PtdIns(4,5)P2) by a phospholipase C to produce inositol trisphosphate (InsP3) and 1,2-diacylglycerol appears to be the initial step in signal transduction for a number of cell-surface interacting stimuli, including thyrotropin-releasing hormone (TRH). In suspensions of membranes isolated from rat pituitary (GH3) cells that were prelabeled to isotopic steady state with [3H]inositol and incubated with ATP, [3H] PtdIns(4,5)P2, and [3H]phosphatidylinositol 4-phosphate, the polyphosphoinositides, and [3H]InsP3 and [3H]inositol bisphosphate, the inositol polyphosphates, accumulated. TRH and GTP stimulated the accumulation of [3H]inositol polyphosphates in time- and concentration-dependent manners; half-maximal effects occurred with 10-30 nM TRH and with 3 microM GTP. A nonhydrolyzable analog of GTP also stimulated [3H] inositol polyphosphate accumulation. Moreover, when TRH and GTP were added together their effects were more than additive. Fixing the free Ca2+ concentration in the incubation buffer at 20 nM, a value below that present in the cytoplasm in vivo did not inhibit stimulation by TRH and GTP of [3H]inositol polyphosphate accumulation. ATP was necessary for basal and stimulated accumulation of [3H]inositol polyphosphates, and a nonhydrolyzable analog of ATP could not substitute for ATP. These data demonstrate that TRH and GTP act synergistically to stimulate the accumulation of InsP3 in suspensions of pituitary membranes and that ATP, most likely acting as substrate for polyphosphoinositide synthesis, was necessary for this effect. These findings suggest that a guanine nucleotide-binding regulatory protein is involved in coupling the TRH receptor to a phospholipase C that hydrolyzes PtdIns(4,5)P2.  相似文献   

11.
Components of the polyphosphoinositide signalling pathway have been identified in stomatal guard cells of Commelina communis L., one of the few plant systems shown unequivocally to be capable of responding to release of inositol 1,4,5-trisphosphate in the cytoplasm by increase in cytoplasmic Ca2+. 'Isolated' epidermal strips of C. communis (in which all cells other than guard cells have been killed by treatment at low pH) were radiolabelled with myo -[2n-3H]inositol or [32P]orthophosphate for 17–18 h. The phosphoinositides and inositol phosphates were extracted. Phosphoinositides were deacylated and the head groups resolved by HPLC. The water-soluble products generated by mild periodate cleavage of HPLC-purified, deacylated lipid fractions were examined. The resulting biochemical analysis led to the identification of: PtdIns, PtdIns3 P , PtdIns4 P , PtdIns(3,4) P 2 and PtdIns(4,5) P 2. Thex inositol phosphates were resolved by HPLC. Preliminary analysis of HPLC-purified putative inositol phosphate fractions resulted in the identification of each inositol phosphate class, that is, Ins P , Ins P 2, Ins P 3, Ins P 4, Ins P 5 and InsP6. Many of these inositol phosphates occurred in different isomeric forms. The presence of 3-phosphorylated phosphoinositides suggests that they may have a role in signalling in stomatal guard cells.  相似文献   

12.
Hydrolysis of the phospholipid phosphatidylinositol 4,5-bisphosphate is thought to be intimately involved in agonist-induced changes in intracellular Ca2+ levels. Recently we have shown that human preovulatory follicular fluid, which induces exocytosis in human sperm, can stimulate a rapid, transient increase in sperm cytosolic [Ca2+] [Thomas & Meizel (1988) Gamete Res. 20, 397-411]. We report here that both a Sephadex G-75 column fraction, derived from follicular fluid, and progesterone (a component of both the G-75 fraction and whole follicular fluid) stimulate rapid hydrolysis of PtdIns(4,5)P2 and PtdIns4P in human sperm. We also report that progesterone stimulates a rapid influx of Ca2+ in human sperm. Human spermatozoa were labelled for 24 h with myo-[3H]inositol and then treated with either the G-75 fraction or progesterone. A 30-65% loss of label was detected in PtdIns(4,5)P2 and PtdIns4P within 15 s of stimulus addition; no changes were observed in PtdIns during 2 min of treatment. The loss of label from both lipids was accompanied by an increase in water-soluble inositol phosphates. Production of both InsP3 and InsP2 was seen within 10 s; however, InsP3 was rapidly removed and had reached control levels by 1 min. Similarly, formation of InsP2 reached a peak by 30 s and then began a decline accompanied by a corresponding increase in InsP. No increases in InsP4 were seen in sperm treated in this fashion. Stimulated hydrolysis of the phosphoinositides and release of inositol phosphates were both blocked by the Ca2+ antagonist La3+. Likewise, the progesterone-induced increase in intracellular Ca2+ was inhibited by La3+, and phosphoinositide hydrolysis stimulated by this hormone was dependent upon the presence of extracellular Ca2+.  相似文献   

13.
Inositol phosphates function as second messengers for a variety of extracellular signals. Ins(1,4,5)P(3) generated by phospholipase C-mediated hydrolysis of phosphatidylinositol bisphosphate, triggers numerous cellular processes by regulating calcium release from internal stores. The Ins(1,4,5)P(3) signal is coupled to a complex metabolic cascade involving a series of phosphatases and kinases. These enzymes generate a range of inositol phosphate derivatives, many of which have signaling roles of their own. We have integrated published biochemical data to build a mass action model for InsP(3) metabolism. The model includes most inositol phosphates that are currently known to interact with each other. We have used this model to study the effects of a G-protein coupled receptor stimulus that activates phospholipase C on the inositol phosphates. We have also monitored how the metabolic cascade interacts with Ins(1,4,5)P(3)-mediated calcium release. We find temporal dynamics of most inositol phosphates to be strongly influenced by the elaborate networking. We also show that Ins(1,3,4,5)P(4) plays a key role in InsP(3) dynamics and allows for paired pulse facilitation of calcium release. Calcium oscillations produce oscillatory responses in parts of the metabolic network and are in turn temporally modulated by the metabolism of InsP(3).  相似文献   

14.
Exposure of A431 human epidermoid carcinoma cells to epidermal growth factor (EGF), bradykinin, and histamine resulted in a time- and concentration-dependent accumulation of the inositol phosphates (InsP) inositol monophosphate, inositol bisphosphate, and inositol trisphosphate (InsP3). Maximal concentrations of EGF (316 ng/ml; approximately 50 nM), bradykinin (1 microM), and histamine (1 mM) resulted in 3-, 6-, and 3-fold increases, respectively, in the amounts of inositol phosphates formed over a 10-min period. The K0.5 values for stimulation were approximately 10 nM, 3 nM, and 10 microM for EGF, bradykinin, and histamine, respectively. EGF and bradykinin stimulated the rapid accumulation of the two isomers of InsP3, Ins(1,3,4)P3, and Ins(1,4,5)P3 as determined by high performance liquid chromatography analysis; maximal accumulation of Ins(1,4,5)P3 occurred within 15 s. EGF and bradykinin also stimulated a rapid (maximal levels attained within 30 s after addition of hormone) and a sustained 4- and 6-fold rise, respectively, in cytosolic free Ca2+ levels as measured by Fura-2 fluorescence. EGF and bradykinin also produced a rapid, although transient, 3- and 5-fold increase, respectively, in cytosolic free Ca2+ after chelation of extracellular Ca2+ with 3 mM EGTA. These data are consistent with the idea that EGF elevates intracellular Ca2+ levels in A431 cells, at least in part, as a result of the rapid formation of Ins(1,4,5)P3 and the consequential release of Ca2+ from intracellular stores.  相似文献   

15.
1-(2-Nitrophenyl)ethyl esters of D-myo-inositol 1,4,5-trisphosphate (InsP3) have been synthesized and shown to have suitable properties for use as photolabile precursors of InsP3. Synthesis was accomplished by treatment of InsP3 with 1-(2-nitrophenyl)diazoethane in a CHCl3/water mixture. This resulted in esterification of each of the three phosphate residues in InsP3, the 1-phosphate being more reactive than the 4- or 5-phosphate. Singly esterified P-1, P-4, and P-5 esters, termed P-1, P-4, and P-5 caged InsP3, were isolated from the reaction mixture by anion-exchange HPLC and characterized by 500-MHz 1H NMR spectroscopy. Each of these caged InsP3 esters exists as a pair of diastereoisomers and was identified by examining the effects of pH and nitrophenyl ring current shielding on the chemical shifts of nonexchangeable inositol protons. 1H NMR spectra of InsP3 were analyzed for comparison. On photolysis the compounds released InsP3 with rate constants of 175 (P-1), 225 (P-4), and 280 s-1 (P-5) as determined by monitoring the aci-nitro decay reaction at pH 7.1, 0.2 M ionic strength, 21 degrees C. Quantum yields determined by steady-state near-UV photolysis were 0.65 +/- 0.08 for each compound. P-4 and P-5 caged InsP3 were the most promising biologically inactive InsP3 precursors since at concentrations up to 50 microM they did not release Ca2+ from smooth muscle sarcoplasmic reticulum (SR) and were not metabolized by vascular smooth muscle InsP3 5-phosphatase or bovine brain InsP3 3-kinase.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
Production of inositol pentakisphosphate in a human T lymphocyte cell line   总被引:1,自引:0,他引:1  
The human T lymphocyte cell line, Jurkat, produced five distinct water soluble, inositol-containing compounds following a period of labeling with 3H-myo-inositol and several hours of incubation in non-radioactive complete medium. The less polar four peaks had been previously shown to be inositol phosphates, InsP through InsP4. Here, we demonstrate that the prominent fifth, very polar, peak was inositol pentakisphosphate, InsP5. The pattern of incorporation of 3H-myo-inositol into InsP5 differed from that of incorporation into other inositol phosphates. InsP5, unlike the second messengers, InsP3 and InsP4, was not increased by perturbation of the T cell receptor/T3 complex.  相似文献   

17.
S Marc  D Leiber    S Harbon 《The Biochemical journal》1988,255(2):705-713
1. In the intact guinea-pig myometrium, carbachol and oxytocin stimulated a specific receptor-mediated phospholipase C activation, catalysing the breakdown of PtdIns(4,5)P2 with the sequential generation of InsP3, InsP2 and InsP. Stimulation of muscarinic receptors also triggered an inhibition of cyclic AMP accumulation caused by prostacyclin. 2. NaF plus AlCl3 mimicked the effects of carbachol and oxytocin by inducing, in a dose-dependent manner, the generation of all three inositol phosphates as well as uterine contractions. AlCl3 enhanced the fluoride effect, supporting the concept that A1F4- was the active species. Under similar conditions, fluoroaluminates activated the guanine nucleotide regulatory protein Gi, reproducing the inhibitory effect of carbachol on cyclic AMP concentrations. 3. Both carbachol- and oxytocin-mediated increases in inositol phosphates, as well as contractions, were insensitive to pertussis toxin, under conditions where the expression of Gi was totally prevented. Cholera toxin, which activates Gs and enhances cyclic AMP accumulation, failed to affect basal or oxytocin-evoked inositol phosphate generation, but induced a slight, though consistent, attenuation of the muscarinic inositol phosphate response, which was similarly evoked by forskolin. 4. The data provide evidence that, in the myometrium, (a) a G protein mediates the generation of inositol phosphates and the Ca2+-dependent contractile event, (b) the relevant G protein that most probably couples muscarinic and oxytocin receptors to phospholipase C is different from Gi and Gs, the proteins that couple receptors to adenylate cyclase, and (c) cyclic AMP does not seem to control the phosphoinositide cycle, but rather exerts a negative regulation at the muscarinic-receptor level.  相似文献   

18.
The formation of inositol phosphates in response to agonists was studied in brain slices, parotid gland fragments and in the insect salivary gland. The tissues were first incubated with [3H]inositol, which was incorporated into the phosphoinositides. All the tissues were found to contain glycerophosphoinositol, inositol 1-phosphate, inositol 1,4-bisphosphate and inositol 1,4,5-trisphosphate, which were identified by using anion-exchange and high-resolution anion-exchange chromatography, high-voltage paper ionophoresis and paper chromatography. There was no evidence for the existence of inositol 1:2-cyclic phosphate. A simple anion-exchange chromatographic method was developed for separating these inositol phosphates for quantitative analysis. Stimulation caused no change in the levels of glycerophosphoinositol in any of the tissues. The most prominent change concerned inositol 1,4-bisphosphate, which increased enormously in the insect salivary gland and parotid gland after stimulation with 5-hydroxytryptamine and carbachol respectively. Carbachol also induced a large increase in the level of inositol 1,4,5-trisphosphate in the parotid. Stimulation of brain slices with carbachol induced modest increase in the bis- and tris-phosphate. In all the tissues studied, there was a significant agonist-dependent increase in the level of inositol 1-phosphate. The latter may be derived from inositol 1,4-bisphosphate, because homogenates of the insect salivary gland contain a bisphosphatase in addition to a trisphosphatase. These results suggest that the earliest event in the stimulus-response pathway is the hydrolysis of polyphosphoinositides by a phosphodiesterase to yield inositol 1,4,5-trisphosphate and inositol 1,4-bisphosphate, which are subsequently hydrolysed to inositol 1-phosphate and inositol. The absence of inositol 1:2-cyclic phosphate could indicate that, at very short times after stimulation, phosphatidylinositol is not catabolized by its specific phosphodiesterase, or that any cyclic derivative liberated is rapidly hydrolysed by inositol 1:2-cyclic phosphate 2-phosphohydrolase.  相似文献   

19.
The InsP3R Ca2+ release channel has a biphasic dependence on cytoplasmic free Ca2+ concentration ([Ca2+]i). InsP3 activates gating primarily by reducing the sensitivity of the channel to inhibition by high [Ca2+]i. To determine if relieving Ca2+ inhibition is sufficient for channel activation, we examined single-channel activities in low [Ca2+]i in the absence of InsP3, by patch clamping isolated Xenopus oocyte nuclei. For both endogenous Xenopus type 1 and recombinant rat type 3 InsP3R channels, spontaneous InsP3-independent channel activities with low open probability Po ( approximately 0.03) were observed in [Ca2+]i < 5 nM with the same frequency as in the presence of InsP3, whereas no activities were observed in 25 nM Ca2+. These results establish the half-maximal inhibitory [Ca2+]i of the channel to be 1.2-4.0 nM in the absence of InsP3, and demonstrate that the channel can be active when all of its ligand-binding sites (including InsP3) are unoccupied. In the simplest allosteric model that fits all observations in nuclear patch-clamp studies of [Ca2+]i and InsP3 regulation of steady-state channel gating behavior of types 1 and 3 InsP3R isoforms, including spontaneous InsP3-independent channel activities, the tetrameric channel can adopt six different conformations, the equilibria among which are controlled by two inhibitory and one activating Ca2+-binding and one InsP3-binding sites in a manner outlined in the Monod-Wyman-Changeux model. InsP3 binding activates gating by affecting the Ca2+ affinities of the high-affinity inhibitory sites in different conformations, transforming it into an activating site. Ca2+ inhibition of InsP3-liganded channels is mediated by an InsP3-independent low-affinity inhibitory site. The model also suggests that besides the ligand-regulated gating mechanism, the channel has a ligand-independent gating mechanism responsible for maximum channel Po being less than unity. The validity of this model was established by its successful quantitative prediction of channel behavior after it had been exposed to ultra-low bath [Ca2+].  相似文献   

20.
F Donié  G Reiser 《FEBS letters》1989,254(1-2):155-158
A membrane preparation from porcine cerebellum displays high-affinity binding sites for [3H]inositol 1,3,4,5-tetrakisphosphate ([3H]InsP4) with a dissociation constant (Kd) of 1.0 nM and a density of 220 fmol/mg protein. Specific binding was maximal in the presence of 25 mM phosphate and at pH 5.0. The receptor site was specific for InsP4, since Ins(1,3,4,5,6)P5 and Ins(1,4,5,6)P4 displaced binding of InsP4 with EC50 values of 0.2 and 0.3 microM, respectively. Ins(1,4,5)P3 and other inositol phosphates were less effective. Using this InsP4 receptor, an assay for measuring tissue content of InsP4 was developed. The detection limit of the assay was 0.1 pmol. In the same tissue samples the amount of Ins(1,4,5)P3 was determined in parallel with a similar assay using a binding protein preparation from beef liver.  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