首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Liver homogenates phosphorylated Ins 1,3,4-P3 to an InsP4 isomer that was distinct from Ins 1,3,4,5-P4. This InsP4 isomer accumulated in vasopressin stimulated hepatocytes prelabeled with myo-[3H]inositol with a time course that lagged behind Ins 1,3,4-P3 formation. The Ins 1,3,4-P3 kinase responsible for its formation was partially purified from rat liver. The enzyme had a Km for Ins 1,3,4-P3 of 0.29 microM, a Km for ATP of 141 microM and was not affected by changes in free Ca2+ in the physiological range. The relationship of this new InsP4 isomer to the inositol phosphate signaling pathway is discussed.  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.
Inositol 1,3,4-trisphosphate (Ins(1,3,4)P3) metabolism has been studied in liver homogenates and in 100,000 x g supernatant and particulate fractions. When liver homogenates were incubated in an "intracellular" medium containing 5 mM MgATP, equal proportions of Ins(1,3,4)P3 were dephosphorylated and phosphorylated. Two inositol tetrakisphosphate (InsP4) products and an inositol pentakisphosphate (InsP5) were detected. The InsP4 isomers were unequivocally identified as inositol 1,3,4,5-tetrakisphosphate (Ins(1,3,4,5)P4) and inositol 1,3,4,6-tetrakisphosphate (Ins(1,3,4,6)P4) by high performance liquid chromatography separation of inositol phosphates, periodate oxidation, alkaline hydrolysis, and stereo-specific polyol dehydrogenase. Ins(1,3,4)P3 5-kinase is a novel enzyme activity and accounted for 16% of the total Ins(1,3,4)P3 phosphorylation. Ins(1,3,4,6)P4 was also shown to be further phosphorylated to inositol 1,3,4,5,6-pentakisphosphate (Ins(1,3,4,5,6)P5) by a kinase not previously known to occur in liver. About 75% of Ins(1,3,4)P3 kinase activities were soluble and were partly purified by anion-exchange fast protein liquid chromatography. The two Ins(1,3,4)P3 kinase activities eluted as a single peak that was well resolved from Ins(1,3,4)P3 phosphatase, Ins(1,3,4,6)P4 5-kinase, and Ins(1,3,4,5)P4 5-phosphatase activities. A further novel observation was that 10 microM Ins(1,3,4,5)P4 inhibited Ins(1,3,4)P3 kinase activities by 60%.  相似文献   

4.
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   

5.
The allosteric effects of various inositol biphosphate (InsP2) isomers and other inositol phosphates, of glycerophosphoinositol phosphates (GroPInsPx) and of phosphoinositides (PtdInsPx) on muscle-type 6-phosphofructo-1-kinase (PFK) were investigated. The binding of these substances to PFK was indirectly estimated by their ability to stabilize the tetrameric enzyme. At near-physiological concentrations of other allosteric effectors, muscle PFK was activated AMP-dependently by Ins(1,4)P2 (Ka = 43 microM), Ins(2,4)P2 (Ka = 70 microM) and GroPIns4P (Ka = 20 microM). These compounds activated PFK by a mechanism similar to that established for activating hexose bisphosphates. Indirect binding experiments indicated minimal Kd,app. values of about 5 microM for the binding of Ins(1,4)P2 in the presence of 0.1 mM-AMP at pH 7.4. This apparent affinity was comparable with that of fructose 1,6-bisphosphate and glucose 1,6-bisphosphate at identical conditions. The enzyme was also found to interact specifically with PtdIns4P (Kd,app. = 37 microM), the inositol phospholipid carrying Ins(1,4)P2 as its head group. The regulatory behaviour of muscle-type PFK in vitro and the concentrations of Ins(1,4)P2 in vivo (between 4 and greater than 50 nmol/g wet wt. of tissue) are consistent with the hypothesis that there is a functional interaction in vivo. Furthermore, a role of PtdIns4P in membrane compartmentation of PFK is suggested. Comparative experiments with liver PFK indicate that these regulatory properties may be relatively specific for the muscle isoform. Unlike muscle PFK, the liver isoform was slightly activated by sub-micromolar concentrations of Ins(1,4,5)P3.  相似文献   

6.
The yeast and Drosophila pathways leading to the production of inositol hexakisphosphate (InsP(6)) have been elucidated recently. The in vivo pathway in humans has been assumed to be similar. Here we show that overexpression of Ins(1,3,4)P(3) 5/6-kinase in human cell lines results in an increase of inositol tetrakisphosphate (InsP(4)) isomers, inositol pentakisphosphate (InsP(5)) and InsP(6), whereas its depletion by RNA interference decreases the amounts of these inositol phosphates. Expression of Ins(1,3,4,6)P(4) 5-kinase does not increase the amount of InsP(5) and InsP(6), although its depletion does block InsP(5) and InsP(6) production, showing that it is necessary for production of InsP(5) and InsP(6). Expression of Ins(1,3,4,5,6)P(5) 2-kinase increases the amount of InsP(6) by depleting the InsP(5) in the cell, and depletion of 2-kinase decreases the amount of InsP(6) and causes an increase in InsP(5). These results are consistent with a pathway that produces InsP(6) through the sequential action of Ins(1,3,4)P(3) 5/6-kinase, Ins(1,3,4,6)P(4) 5-kinase, and Ins(1,3,4,5,6)P5 2-kinase to convert Ins(1,3,4)P(3) to InsP(6). Furthermore, the evidence implicates 5/6-kinase as the rate-limiting enzyme in this pathway.  相似文献   

7.
1. A screen for agonists capable of stimulating the formation of inositol phosphates in erythrocytes from 5-day-old chickens revealed the presence of a population of phosphoinositidase C-linked purinergic receptors. 2. If chicken erythrocytes prelabelled with [3H]Ins were exposed to a maximal effective dose of adenosine 5'-[beta-thio]diphosphate for 30 s, the agonist-stimulated increment in total [3H]inositol phosphates was confined to [3H]Ins(1,4,5)P3, Ins(1,3,4,5)P4 and InsP2. After 40 min stimulation, the radiolabelling of nearly all of the [3H]inositol phosphates that have been detected in these extracts [Stephens, Hawkins & Downes (1989) Biochem. J. 262, 727-737] had risen. However, some of these increases [especially those in Ins(3,4,5,6)P4 and Ins(1,3,4,5,6)P5] were accountable for almost entirely by increases in specific radioactivity rather than in mass. 3. The effect of purinergic stimulation on the rate of incorporation of [32P]Pi in the medium into the gamma-phosphate group of ATP and InsP4 and InsP5 was also measured. After 40 min stimulation, the incorporation of 32P into Ins(1,3,4,6)P4, Ins(1,3,4,5)P4, Ins(3,4,5,6)P4 and Ins(1,3,4,5,6)P5 was significantly elevated, whereas the mass of the last two and the specific radioactivity of the gamma-phosphate of ATP were unchanged compared with control erythrocyte suspensions. 4. In control suspensions of avian erythrocytes, the specific radioactivity of the individual phosphate moieties of Ins(1,3,4,6)P4 increased through the series 1, 6, 4 and 3 [Stephens & Downes (1990) Biochem. J. 265, 435-452]. This pattern of 32P incorporation is not the anticipated outcome of 6-hydroxy phosphorylation of Ins(1,3,4)P3 [the assumed route of synthesis of Ins(1,3,4,6)P4]. Although adenosine [beta-thio]diphosphate significantly stimulated the accumulation of [3H]Ins(1,3,4)P3, and despite the fact that avian erythrocyte lysates were shown to possess a chromatographically distinct, soluble, ATP-dependent, Ins(1,3,4)P3 6-hydroxykinase activity, purinergic stimulation of intact cells did not significantly alter the pattern of incorporation of [32P]Pi into the individual phosphate moieties of Ins(1,3,4,6)P4. These results suggest that the route of synthesis of this inositol phosphate species is not changed during the presence of an agonist.  相似文献   

8.
The spectrum of inositol phosphate isomers present in avian erythrocytes was investigated in qualitative and quantitative terms. Inositol phosphates were isolated in micromolar quantities from turkey blood by anion-exchange chromatography on Q-Sepharose and subjected to proton n.m.r. and h.p.l.c. analysis. We employed a h.p.l.c. technique with a novel, recently described complexometric post-column detection system, called 'metal-dye detection' [Mayr (1988) Biochem. J. 254, 585-591], which enabled us to identify non-radioactively labelled inositol phosphate isomers and to determine their masses. The results indicate that avian erythrocytes contain the same inositol phosphate isomers as mammalian cells. Denoted by the 'lowest-locant rule' [NC-IUB Recommendations (1988) Biochem. J. 258, 1-2] irrespective of true enantiomerism, these are Ins(1,4)P2, Ins(1,6)P2, Ins(1,3,4)P3, Ins(1,4,5)P3, Ins(1,3,4,5)P4, Ins(1,3,4,6)P4, Ins(1,4,5,6)P4, Ins(1,3,4,5,6)P5, and InsP6. Furthermore, we identified two inositol trisphosphate isomers hitherto not described for mammalian cells, namely Ins(1,5,6)P3 and Ins(2,4,5)P3. The possible position of these two isomers in inositol phosphate metabolism and implications resulting from absolute abundances of inositol phosphates are discussed.  相似文献   

9.
Stimulation of the human T-lymphocyte cell line Jurkat via the T-cell receptor/CD3 complex by an anti-CD3 antibody (OKT3) induced time-dependent changes in the intracellular concentrations of multiple inositol polyphosphate (InsPn) isomers. Quantitative mass analysis by anion-exchange HPLC and a recently developed postcolumn dye system (Mayr, G. W. (1988) Biochem. J. 254, 585-591) revealed basal intracellular concentrations between less than 5 pmol/10(9) cells for Ins(1,3,4,5)P4 and 6380 +/- 355 pmol/10(9) cells for InsP6. Time course analysis of samples from stimulated Jurkat T-cells showed an increase of Ins(1,3,4,5)P4 to 1125 +/- 125 pmol/10(9) cells within 10 min and remained elevated over more than 30 min. Moreover, increases of the intracellular concentrations of Ins(1,3,4,6)P4, Ins(1,4,5,6)P4, and/or Ins(3,4,5,6)P4 (determined as the enantiomeric mixture), Ins(1,3,4,5,6)P5, Ins(1,2,3,4,6)P5 and InsP6 were observed. In contrast, the concentration of Ins(1,2,4,5,6)P5 and/or Ins(2,3,4,5,6)P5 (determined as the enantiomeric mixture) decreased after stimulation. Using cytosolic extracts from Jurkat T-lymphocytes incubated with purified Ins(1,3,4,5,6)P5, Ins(1,2,3,4,6)P5, or Ins(1,2,4,5,6)P5/Ins(2,3,4,5,6)P5 three enzyme activities were observed. Ins(1,3,4,5,6)P5 was dephosphorylated by a phosphatase removing a phosphate group at the 1 and/or 3 position resulting in the formation of Ins(1,4,5,6)P4 and/or Ins(3,4,5,6)P4 (determined as the enantiomeric mixture). Ins(1,2,3,4,6)P5 was metabolized by a specific phosphatase that cleaved the phosphate group at the 2 position, thereby generating the product Ins(1,3,4,6)P4. On the other hand, Ins(1,2,4,5,6)P5/Ins(2,3,4,5,6)P5 was phosphorylated by a 1/3-kinase activity to InsP6. Together novel receptor-mediated metabolic pathways of inositol polyphosphates were demonstrated in human T-lymphocytes, and corresponding enzyme activities for the inositol pentakisphosphate metabolism were found in cell lysates.  相似文献   

10.
Stimulation of NIH-3T3 cells with prostaglandin F2 alpha (PGF2 alpha) caused a dose- and time-dependent generation of inositol phosphates. The first detectable changes were in the levels of Ins(1,4,5)P3 and Ins(1,3,4,5)P4. Increases in Ins(1,3,4)P3, InsP2 and InsP were detected later, and only minor changes were observed in putative InsP5 or InsP6. The accumulation of inositol phosphates was synergistically increased by the addition of calf serum, whereas PGF2 alpha had no effects on cell proliferation in either the presence or the absence of calf serum. Stimulation of a different clone of NIH-3T3 cells (AmNIH-3T3) or Swiss 3T3 cells with PGF2 alpha resulted in both inositol phospholipid breakdown and cell proliferation. No differences were found in the characteristics of PGF2 alpha-stimulated inositol phosphate generation between the two clones of NIH-3T3 cells, nor was there any difference in receptor number of Kd. These results question the role of inositol phospholipid breakdown in mitogenesis and demonstrate significant differences in the biochemical properties of apparently the 'same' cells.  相似文献   

11.
An enzyme which catalyses the ATP-dependent phosphorylation of inositol 1,4,5-trisphosphate [Ins(1,4,5)P3] was purified approx. 180-fold from rat brain cytosol by (NH4)2SO4 precipitation, chromatography through hydroxyapatite, anion-exchange fast protein liquid chromatography and gel-filtration chromatography. Gel filtration on Sepharose 4B CL gives an Mr of 200 x 10(3) for the native enzyme. The inositol tetrakisphosphate (InsP4) produced by the enzyme has the chromatographic, chemical and metabolic properties of Ins(1,3,4,5)P4. Ins(1,4,5)P3 3-kinase displays simple Michaelis-Menten kinetics for both its substrates, having Km values of 460 microM and 0.44 microM for ATP and Ins(1,4,5)P3 respectively. When many of the inositol phosphates known to occur in cells were tested, only Ins(1,4,5)P3 was a substrate for the enzyme; the 2,4,5-trisphosphate was not phosphorylated. Inositol 4,5-bisphosphate and glycerophosphoinositol 4,5-bisphosphate were phosphorylated much more slowly than Ins(1,4,5)P3. CTP, GTP and adenosine 5'-[gamma-thio]triphosphate were unable to substitute for ATP. When assayed under conditions of first-order kinetics, Ins(1,4,5)P3 kinase activity decreased by about 40% as the [Ca2+] was increased over the physiologically relevant range. This effect was insensitive to the presence of calmodulin and appeared to be the result of an increase in the Km of the enzyme for Ins(1,4,5)P3. Preincubation with ATP and the purified catalytic subunit of cyclic AMP-dependent protein kinase did not affect the rate of phosphorylation of Ins(1,4,5)P3 when the enzyme was assayed at saturating concentrations of Ins(1,4,5)P3 or at concentrations close to its Km for this substrate.  相似文献   

12.
A simple procedure for assay of Ins(1,4,5)P3 5-phosphatase is described. The reaction products [( 3H]Ins(1,4)P2, [3H]InsP and myo-[3H]inositol) are completely separated from one another, with quantitative yield, on Amprep SAX (100 mg) minicolumns. [3H]Ins(1,4,5)P3 [and [3H]Ins(1,3,4,5)P4] are adsorbed to the columns but not released to any appreciable extent by the elution conditions used. In GH3 cells, the stepwise dephosphorylation of [3H]Ins(1,4,5)P3 to myo-[3H]inositol was demonstrated, and was inhibited by 2.3-bisphosphoglycerate. The Km of the soluble form of the enzyme was lower in GH3 cells (8-13 microM) than in IMR-32 cells (26-32 microM) or in rat cerebral-cortical samples (22 microM. The Km of the particulate form of the enzyme was similar in all three preparations (10-16 microM). The pH profiles of the two soluble 5-phosphatases differed, with a wider pH optimum for the GH3-cell activity than for the IMR-32-cell activity. The soluble and particulate GH3 enzymes were more sensitive than the corresponding IMR-32 enzymes to inhibition by p-hydroxymercuribenzoate, whereas there were no differences in their sensitivities to glucose 6-phosphate, 2,3-bisphosphoglycerate, fructose 1.6- and 2.6-bisphosphate and non-radioactive Ins(1,3,4,5)P4. Dialysis of the soluble fractions and washing of the particulate fractions did not affect the inhibitor sensitivities, except for the soluble IMR-32 fraction and p-hydroxymercuribenzoate. The Km value of the soluble GH3 5-phosphatase activity was lower, and the inhibition by Ins(1,3,4,5)P4 greater, after adsorption to and elution from phosphocellulose. It is concluded that there are qualitative differences in the properties of the soluble 5-phosphatase activity from GH3 and IMR-32 cells.  相似文献   

13.
When [3H]inositol-prelabelled N1E-115 cells were stimulated with carbamylcholine (CCh) (100 microM), high K+ (60 mM), and prostaglandin E1 (PGE1) (10 microM), a transient increase in [3H]inositol pentakisphosphate (InsP5) accumulation was observed. The accumulation reached its maximum level at 15 s and had declined to the basal level at 2 min. CCh, high K+, and PGE1 also caused accumulations of [3H]inositol 1,4,5-trisphosphate [Ins(1,4,5)P3], [3H]inositol 1,3,4,6-tetrakisphosphate [Ins(1,3,4,6)P4], and [3H]inositol hexakisphosphate (InsP6). Muscarine and CCh induced accumulations of [3H]Ins(1,4,5)P3, [3H]-Ins(1,3,4,6)P4, [3H]InsP5, and [3H]InsP6 with a similar potency and exerted these maximal effects at 100 microM, whereas nicotine failed to do so at 1 mM. With a slower time course, CCh, high K+, and PGE1 caused accumulations of [3H]-inositol 1,3,4-trisphosphate [Ins(1,3,4)P3] and [3H]inositol 1,3,4,5-tetrakisphosphate [Ins(1,3,4,5)P4]. In an N1E-115 cell homogenate, [3H]Ins(1,4,5)P3, [3H]Ins(1,3,4,5)P4, and [3H]Ins(1,3,4)P3 were converted to [3H]InsP5 through [3H]-Ins(1,3,4,6)P4. The above results indicate that Ins(1,3,4,6)P4, InsP5, and InsP6 are rapidly formed by several kinds of stimulants in N1E-115 cells.  相似文献   

14.
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   

15.
1. Basal and carbachol-stimulated accumulations of isomeric [3H]inositol mono-, bis-, tris- and tetrakis-phosphates were examined in rat cerebral-cortex slices labelled with myo-[2-3H]inositol. 2. In control samples the major [3H]inositol phosphates detected were co-eluted on h.p.l.c. with Ins(1)P, Ins(4)P (inositol 1- and 4-monophosphate respectively), Ins(1,4)P2 (inositol 1,4-bisphosphate), Ins(1,4,5)P3 (inositol 1,4,5-tris-phosphate) and Ins(1,3,4,5)P4 (inositol 1,3,4,5-tetrakisphosphate). 3. After stimulation to steady state with carbachol, accumulation of each of these products was markedly increased. 4. Agonist stimulation, however, also evoked much more dramatic increased accumulations of a second [3H]inositol trisphosphate, which was co-eluted on h.p.l.c. with authentic Ins(1,3,4)P3 (inositol 1,3,4-trisphosphate) and of three further [3H]inositol bisphosphates ([3H]InsP2(s]. 5. Examination of the latter by chemical degradation by periodate oxidation and/or h.p.l.c. allowed identification of these as [3H]Ins(1,3)P2, [3H]Ins(3,4)P2 and [3H]Ins(4,5)P2 (inositol 1,3-, 3,4- and 4,5-bisphosphates respectively), which respectively accounted for about 22%, 8% and 3% of total [3H]InsP2 in extracts from stimulated tissue slices. 6. By using a h.p.l.c. method which clearly resolves Ins(1,3,4,5)P4 and Ins(1,3,4,6)P4 (inositol 1,3,4,6-tetrakisphosphate), only the former isomer could be detected in extracts from either control or stimulated tissue slices. Similarly, [3H]inositol pentakis- and hexakis-phosphates were not detectable either in the presence or absence of carbachol under the radiolabelling conditions described. 7. The catabolism of [3H]Ins(1,4,5)P3 and [3H]Ins(1,3,4)P3 by cell-free preparations from cerebral cortex was also studied. 8. In the presence of Mg2+, [3H]Ins(1,4,5)P3 was specifically dephosphorylated via [3H]Ins(1,4)P2 and [3H]Ins(4)P to free [3H]inositol, whereas [3H]Ins(1,3,4)P3 was degraded via [3H]Ins(3,4)P2 and, to a lesser extent, via [3H]Ins(1,3)P2 to D- and/or L-[3H]Ins(1)P and [3H]inositol. 9. In the presence of EDTA, hydrolysis of [3H]Ins(1,4,5)P3 was greater than or equal to 95% inhibited, whereas [3H]Ins(1,3,4)P3 was still degraded, but yielded only a single [3H]InsP2 identified as [3H]Ins(1,3)P2. 10. The significance of these observations with cell-free preparations is discussed in relation to the proportions of the separate isomeric [3H]inositol phosphates measured in stimulated tissue slices.  相似文献   

16.
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   

17.
We have identified, isolated, and characterized a second inositol polyphosphate-5-phosphatase enzyme from the soluble fraction of human platelets. The enzyme hydrolyzes inositol 1,4,5-trisphosphate (Ins (1,4,5)P3) to inositol 1,4-bisphosphate (Ins(1,4)P2) with an apparent Km of 24 microM and a Vmax of 25 mumol of Ins(1,4,5)P3 hydrolyzed/min/mg of protein. The enzyme hydrolyzes inositol (1,3,4,5)-tetrakisphosphate (Ins(1,3,4,5)P4) at a rate of 1.3 mumol of Ins(1,3,4,5)P4 hydrolyzed/min/mg of protein with an apparent Km of 7.5 microM. The enzyme also hydrolyzes inositol 1,2-cyclic 4,5-trisphosphate (cIns(1:2,4,5)P3) and Ins(4,5)P2. We purified this enzyme 2,200-fold from human platelets. The enzyme has a molecular mass of 75,000 as determined by both sodium dodecyl sulfate-polyacrylamide gel electrophoresis and by gel filtration chromatography. The enzyme requires magnesium ions for activity and is not inhibited by calcium ions. The 75-kDa inositol polyphosphate-5-phosphatase enzyme differs from the previously identified platelet inositol polyphosphate-5-phosphatase as follows: molecular size (75 kDa versus 45 kDa), affinity for Ins(1,3,4,5)P4 (Km 7.5 microM versus 0.5 microM), Km for Ins(1,4,5)P3 (24 microM versus 7.5 microM), regulation by protein kinase C, wherein the 45-kDa enzyme is phosphorylated and activated while the 75-kDa enzyme is not. The 75-kDa enzyme is inhibited by lower concentrations of phosphate (IC50 2 mM versus 16 mM for the 45-kDa enzyme) and is less inhibited by Ins(1,4)P2 than is the 45-kDa enzyme. The levels of inositol phosphates that act in calcium signalling are likely to be regulated by the interplay of these two enzymes both found in the same cell.  相似文献   

18.
Anion-exchange h.p.l.c. analysis of [3H]inositol phosphates derived from glucose-stimulated isolated pancreatic islets that had been prelabelled with myo-[3H]inositol revealed that the predominant inositol trisphosphate was the 1,3,4-isomer [Ins(1,3,4)P3]. The 1,4,5-isomer [Ins(1,4,5)P3] was also detectable, as was a more polar inositol phosphate with the chromatographic properties of inositol 1,3,4,5-tetrakisphosphate [Ins(1,3,4,5)P4]. Glucose-induced accumulation of Ins(1,3,4)P3 was augmented by Li+ and occurred after maximal accumulation of Ins(1,4,5)P3. These findings suggest a possible role for Ins(1,3,4)P3 or its probable precursor Ins(1,3,4,5)P4 in stimulus-secretion coupling in pancreatic islets.  相似文献   

19.
Phospholipase C cleaves phosphatidylinositol 4,5-bisphosphate to form both inositol 1,4,5-trisphosphate (Ins(1,4,5)P3) and inositol 1,2-cyclic 4,5-trisphosphate (cInsP3). The further metabolism of these inositol trisphosphates is determined by two enzymes: a 3-kinase and a 5-phosphomonoesterase. The first enzyme converts Ins(1,4,5)P3 to inositol 1,3,4,5-tetrakisphosphate (InsP4), while the latter forms inositol 1,4-bisphosphate and inositol 1,2-cyclic 4-bisphosphate from Ins(1,4,5)P3 and cInsP3, respectively. The current studies show that the 3-kinase is unable to phosphorylate cInsP3. Also, the 5-phosphomonoesterase hydrolyzes InsP4 with an apparent Km of 0.5-1.0 microM to form inositol 1,3,4-trisphosphate at a maximal velocity approximately 1/30 that for Ins(1,4,5)P3. The apparent affinity of the enzyme for the three substrates is InsP4 greater than Ins(1,4,5)P3 greater than cInsP3; however, the rate at which the phosphatase hydrolyzes these substrates is Ins(1,4,5)P3 greater than cInsP3 greater than InsP4. The 5-phosphomonoesterase and 3-kinase enzymes may control the levels of inositol trisphosphates in stimulated cells. The 3-kinase has a low apparent Km for Ins(1,4,5)P3 as does the 5-phosphomonoesterase for InsP4, implying that the formation and breakdown of InsP4 may proceed when both it and its precursor are present at low levels. Ins(1,4,5)P3 is utilized by both the 3-kinase and 5-phosphomonoesterase, while cInsP3 is utilized relatively poorly only by the 5-phosphomonoesterase. These findings imply that inositol cyclic trisphosphate may be metabolized slowly after its formation in stimulated cells.  相似文献   

20.
The effects of Li+ on carbachol-stimulated phosphoinositide metabolism were examined in rat cerebral-cortex slices labelled with myo-[2-3H]inositol. The muscarinic agonist carbachol evoked an enhanced steady-state accumulation of [3H]inositol monophosphate ([3H]InsP1), [3H]inositol bisphosphate ([3H]InsP2), [3H]inositol 1,3,4-trisphosphate ([3H]Ins(1,3,4)P3), [3H]inositol 1,4,5-trisphosphate ([3H]Ins(1,4,5)P3) and [3H]inositol tetrakisphosphate ([3H]InsP4). Li+ (5 mM), after a 10 min lag, severely attenuated carbachol-stimulated [3H]InsP4 accumulation while simultaneously potentiating accumulation of both [3H]InsP1 and [3H]InsP2 and, at least initially, of [3H]Ins(1,3,4)P3. These data are consistent with inhibition of inositol mono-, bis- and 1,3,4-tris-phosphate phosphatases to different degrees by Li+ in brain, but are not considered to be completely accounted for in this way. Potential direct and indirect mechanisms of the inhibitory action of Li+ on [3H]InsP4 accumulation are considered. The present results stress the complex action of Li+ on cerebral inositol metabolism and indicate that more complex mechanisms than are yet evident may regulate this process.  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