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 22 毫秒
1.
Antigen-mediated exocytosis in intact rat basophilic leukemia (RBL-2H3) cells is associated with substantial hydrolysis of membrane inositol phospholipids and an elevation in concentration of cytosol Ca2+ ([ Ca2+i]). Paradoxically, these two responses are largely dependent on external Ca2+. We report here that cells labeled with myo-[3H]inositol and permeabilized with streptolysin O do release [3H]inositol 1,4,5-trisphosphate upon stimulation with antigen or guanosine 5'-O-(3-thiotriphosphate) (GTP gamma S) at low (less than 100 nM) concentrations of free Ca2+. The response, however, is amplified by increasing free Ca2+ to 1 microM. The subsequent conversion of the trisphosphate to inositol 1,3,4,5-tetrakisphosphate is enhanced also by the increase in free Ca2+. Although [3H]inositol 1,4,5-trisphosphate accumulates in greater amounts than is the case in intact cells, [3H]inositol 1,4-bisphosphate is still the major product in permeabilized cells even when the further metabolism of [3H]inositol 1,4,5-trisphosphate is suppressed (by 77%) by the addition of excess (1000 microM) unlabeled inositol 1,4,5-trisphosphate and the phosphatase inhibitor 2,3-bisphosphoglycerate. It would appear that either the activity of the membrane 5-phosphomonoesterase allows virtually instantaneous dephosphorylation of the inositol 1,4,5-trisphosphate under all conditions tested or both phosphatidylinositol 4-monophosphate and the 4,5-bisphosphate are substrates for the activated phospholipase C. The latter alternative is supported by the finding that permeabilized cells, which respond much more vigorously to high (supraoptimal) concentrations of antigen than do intact RBL-2H3 cells, produce substantial amounts of [3H]inositol 1,4-bisphosphate before any detectable increase in levels of [3H]inositol 1,4,5-trisphosphate.  相似文献   

2.
Lysed mouse thymocytes release [3H]inositol 1,4,5 trisphosphate from [3H]inositol-labelled phosphatidyl inositol 4,5-bisphosphate in response to GTP gamma S, and rapidly phosphorylate [3H]inositol 1,4,5-trisphosphate to [3H]inositol 1,3,4,5-tetrakisphosphate. The rate of phosphorylation is increased approximately 7-fold when the free [Ca2+] in the lysate is increased from 0.1 to 1 microM, the range in which the cytosolic free [Ca2+] increases in intact thymocytes in response to the mitogen concanavalin A. Stimulation of the intact cells with concanavalin A also results in a rapid and sustained increase in the amount of inositol 1,3,4,5-tetrakisphosphate, and a much smaller transient increase in 1,4,5-trisphosphate. Lowering [Ca2+] in the medium from 0.4 mM to 0.1 microM before addition of concanavalin A reduces accumulation of inositol 1,3,4,5-tetrakisphosphate by at least 3-fold whereas the increase in inositol 1,4,5-trisphosphate is sustained rather than transient. The data imply that in normal medium the activity of the inositol 1,4,5-trisphosphate kinase increases substantially in response to the rise in cytosolic free [Ca2+] generated by concanavalin A, accounting for both the transient accumulation of inositol 1,4,5-trisphosphate and the sustained high levels of inositol 1,3,4,5-tetrakisphosphate. Inositol 1,3,4,5-tetrakisphosphate is a strong candidate for the second messenger for Ca2+ entry across the plasma membrane. This would imply that the inositol polyphosphates regulate both Ca2+ entry and intracellular Ca2+ release, with feedback control of the inositol polyphosphate levels by Ca2+.  相似文献   

3.
The role of Ca2+ in the generation of inositol phosphates was investigated using rat pancreatic islets after steady state labeling with myo-[2-3H]inositol. Depolarizing K+ concentrations (24 mM) evoked early (2 s) increases in inositol 1,4,5-trisphosphate (Ins-1,4,5-P3) and inositol 1,3,4,5-tetrakisphosphate (Ins-1,3,4,5-P4) as measured by high performance anion-exchange chromatography. The increase in Ins-1,4,5-P3 was transient and was followed by a more pronounced rise in Ins-1,3,4-P3. These effects were dependent on the presence of extracellular Ca2+ but were not secondary to release of either neurotransmitters or metabolites of arachidonic acid. K+ also promoted the breakdown of phosphatidylinositol 4,5-bisphosphate (PtdIns-4,5-P2) and of the other phosphoinositides. Glucose (16.7 mM) was less marked in its effects but still promoted rapid increases in Ins-1,3,4,5-P4 (2 s) and Ins-1,4,5-P3 (10 s) and a slower rise in Ins-1,3,4-P3 (30 s). The levels of all three metabolites rose steadily over 10 min stimulation. These responses to glucose could be largely, although not entirely, inhibited by depletion of extracellular Ca2+ or by Ca2+ channel blockade with verapamil (20 microM). Carbamylcholine (0.5 mM) was the most potent stimulus used evoking early rises in Ins-1,4,5-P3 and Ins-1,3,4,5-P4 (2 s) followed by Ins-1,3,4-P3 (10 s), effects which were only partially dependent on extracellular Ca2+. The results suggest that a Ca2+-mediated PtdIns-4,5-P2 hydrolysis accounts for most of the Ins-1,4,5-P3 generated in response to glucose but not carbamylcholine. In addition, glucose may exert effects on inositol phosphate metabolism which are Ca2+ independent.  相似文献   

4.
myo-Inositol 1,4,5-trisphosphate is an intracellular second messenger generated from the hydrolysis of phosphatidylinositol 4,5-bisphosphate by phospholipase C. In the present study, we have used the abilities of inositol 1,4,5-trisphosphate to inhibit inositol 1,4,5-tris[32P]phosphate binding and to stimulate release of sequestered stores of 45Ca2+ to assay the mass of inositol 1,4,5-trisphosphate in extracts derived from [3H]inositol-prelabeled chemoattractant-stimulated neutrophils. These assays are specific for inositol 1,4,5-trisphosphate since the relative capacity of the extracts to compete with inositol 1,4,5-tris[32P]phosphate binding and to release 45Ca2+ correlated well with the [3H]inositol 1,4,5-trisphosphate content of the extract as determined by high pressure liquid chromatography. No correlation of these activities was observed with the content in the extract of either [3H]inositol 1,3,4-trisphosphate or [3H]inositol 1,3,4,5-tetrakisphosphate, whose formation exhibited kinetics distinct from [3H]inositol 1,4,5-trisphosphate. Thus, within 10 s of stimulation with 10 nM formyl-methionyl-leucyl-phenylalanine, the inositol 1,4,5-trisphosphate content of the extract increased from 0.05 to 0.55 pmol/10(6) cells, equivalent to a change in intracellular concentration from 100 nM to 1.1 microM. These studies demonstrate that neutrophils produce sufficient quantities of inositol 1,4,5-trisphosphate to mobilize Ca2+ from intracellular stores.  相似文献   

5.
The metabolism of the calcium mobilizing inositol-1,4,5-trisphosphate (IP3) isomer was studied in myo-[3H]inositol labeled, chemoattractant-stimulated human polymorphonuclear neutrophils (PMNs), and in PMN lysates. It was determined that 1,4,5-IP3 is metabolized in vitro by two distinct pathways: 1) by sequential dephosphorylation to 1,4-IP2, 4-IP1, and inositol or 2) by ATP dependent conversion to 1,3,4,5-IP4, followed by dephosphorylation to form 1,3,4-IP3, 3,4-IP2, 3-IP1, and inositol. In PMNs stimulated with 0.1 microM N-formyl-methionyl-leucyl-phenylalanine (fMet-Leu-Phe), 1,4-IP2, 1,4,5-IP3, and IP4, were elevated by 5 s; whereas production of 1,3,4-IP3, 3,4-IP2, and IP1 occurred only after an initial lag (approximately 15 s). The predominant IP1 isomer formed in fMet-Leu-Phe-stimulated cells was 4-IP1. Production of 1,3,4-IP3 and 3,4-IP2 was markedly reduced (17 and 35% of control, respectively) in fMet-Leu-Phe-stimulated cells pretreated to prevent a rise in intracellular calcium ([Ca2+]i). PMNs were also stimulated with leukotriene B4 (LTB4) since this agent is a poor activator of the respiratory burst compared to fMet-Leu-Phe. Peak levels (5 s) of 1,4,5-IP3 were equivalent after stimulation with 0.1 microM fMet-Leu-Phe versus 0.1 microM LTB4 (320 +/- 38% versus 378 +/- 38% of control values, respectively; n = 5); however, at 30 s, 1,4,5-IP3 remained elevated only in fMet-Leu-Phe-stimulated cells. Similarly, elevation of [Ca2+]i was more prolonged in response to 0.1 microM fMet-Leu-Phe (greater than 3 min) versus LTB4 (1 min). Thus, signal transduction in PMNs may be modulated by both the duration of the initial 1,4,5-IP3 signal and by the metabolic pathway(s) utilized to convert this IP3 isomer to other, potentially active inositol phosphate products.  相似文献   

6.
The analysis of the inositol cycle in Dictyostelium discoideum cells is complicated by the limited uptake of [3H]inositol (0.2% of the applied radioactivity in 6 h), and by the conversion of [3H]inositol into water-soluble inositol metabolites that are eluted near the position of inositol 1,4,5-trisphosphate [Ins(1,4,5)P3] on anion-exchange h.p.l.c. columns. The uptake was improved to 2.5% by electroporation of cells in the presence of [3H]inositol; electroporation was optimal at two 210 microseconds pulses of 7 kV. Cells remained viable and responsive to chemotactic signals after electroporation. The intracellular [3H]inositol was rapidly metabolized to phosphatidylinositol and more slowly to phosphatidylinositol phosphate and phosphatidylinositol bisphosphate. More than 85% of the radioactivity in the water-soluble extract that was eluted on Dowex columns as Ins(1,4,5)P3 did not co-elute with authentic [32P]Ins(1,4,5)P3 on h.p.l.c. columns. Chromatography of the extract by ion-pair reversed-phase h.p.l.c. provided a good separation of the polar inositol polyphosphates. Cellular [3H]Ins(1,4,5)P3 was identified by (a) co-elution with authentic [32P]Ins(1,4,5)P3 and (b) degradation by a partially purified Ins(1,4,5)P3 5-phosphatase from rat brain. The chemoattractant cyclic AMP and the non-hydrolysable analogue guanosine 5'-[gamma-thio]triphosphate induced a transient accumulation of radioactivity in Ins(1,4,5)P3; we did not detect radioactivity in inositol 1,3,4-trisphosphate or inositol 1,3,4,5-tetrakisphosphate [Ins(1,3,4,5)P4]. In vitro, Ins(1,4,5)P3 was metabolized to inositol 1,4- and 4,5-bisphosphate, but not to Ins(1,3,4,5)P4 or another tetrakisphosphate isomer. We conclude that Dictyostelium has a receptor- and G-protein-stimulated inositol cycle which is basically identical with that in mammalian cells, but the metabolism of Ins(1,4,5)P3 is probably different.  相似文献   

7.
Addition of 1 mM-carbachol to [3H]inositol-labelled rat parotid slices stimulated rapid formation of [3H]inositol 1,3,4,5-tetrakisphosphate, the accumulation of which reached a peak 20 s after stimulation, and then declined rapidly towards a new steady state. The initial rate of formation of inositol 1,3,4,5-tetrakisphosphate was slower than that for inositol 1,4,5-trisphosphate. The radioactivity in [3H]inositol 1,3,4,5-tetrakisphosphate fell quickly in carbachol-stimulated and then atropine-blocked parotid slices, suggesting that it is rapidly metabolized during stimulation. Parotid homogenates rapidly dephosphorylated inositol 1,4,5-trisphosphate, inositol 1,3,4,5-tetrakisphosphate and, less rapidly, inositol 1,3,4-trisphosphate. Inositol 1,3,4,5-tetrakisphosphate was specifically hydrolysed to a compound with the chromatographic properties of inositol 1,3,4-trisphosphate. The only 3H-labelled phospholipids that we could detect in parotid slices labelled with [3H]inositol for 90 min were phosphatidylinositol, phosphatidylinositol 4-phosphate and phosphatidylinositol 4,5-bisphosphate. Parotid homogenates synthesized inositol tetrakisphosphate from inositol 1,4,5-trisphosphate. This activity was dependent on the presence of ATP. We suggest that, during carbachol stimulation of parotid slices, the key event in inositol lipid metabolism is the activation of phosphatidylinositol 4,5-bisphosphate-specific phospholipase C. The inositol 1,4,5-trisphosphate thus liberated is metabolized in two distinct ways; by direct hydrolysis of the 5-phosphate to form inositol 1,4-bisphosphate and by phosphorylation to form inositol 1,3,4,5-tetrakisphosphate and hence, by hydrolysis of this tetrakisphosphate, to form inositol 1,3,4-trisphosphate.  相似文献   

8.
Formation and metabolism of inositol 1,3,4,5-tetrakisphosphate in liver   总被引:29,自引:0,他引:29  
The inositol lipid pools of isolated rat hepatocytes were labeled with [3H]myo-inositol, stimulated maximally with vasopressin and the relative contents of [3H]inositol phosphates were measured by high performance liquid chromatography. Inositol 1,4,5-trisphosphate accumulated rapidly (peak 20 s), while inositol 1,3,4-trisphosphate and a novel inositol phosphate (ascribed to inositol 1,3,4,5-tetrakisphosphate) accumulated at a slower rate over 2 min. Incubation of hepatocytes with 10 mM Li+ prior to vasopressin addition selectively augmented the levels of inositol monophosphate, inositol 1,4-bisphosphate, and inositol 1,3,4-trisphosphate. A kinase was partially purified from liver and brain cortex which catalyzed an ATP-dependent phosphorylation of [3H]inositol 1,4,5-trisphosphate to inositol 1,3,4,5-tetrakisphosphate. Incubation of purified [3H]inositol 1,3,4,5-tetrakisphosphate with diluted liver homogenate produced initially inositol 1,3,4-trisphosphate and subsequently inositol 1,3-bisphosphate, the formation of which could be inhibited by Li+. The data demonstrate that the most probable pathway for the formation of inositol 1,3,4,5-tetrakisphosphate is by 3-phosphorylation of inositol 1,4,5-trisphosphate by a soluble mammalian kinase. Degradation of both compounds occurs first by a Li+-insensitive 5-phosphatase and subsequently by a Li+-sensitive 4-phosphatase. The prolonged accumulation of both inositol 1,4,5-trisphosphate and inositol 1,3,4,5-tetrakisphosphate in vasopressin-stimulated hepatocytes suggest that they have separate second messenger roles, perhaps both relating to Ca2+-signalling events.  相似文献   

9.
Previous studies with antigen-stimulated rat basophilic leukemia (RBL-2H3) cells indicated the formation of multiple isomers of each of the various categories of inositol phosphates. The identities of the different isomers have been elucidated by selective labeling of [3H]inositol 1,3,4,5-tetrakisphosphate with [32P]phosphate in the 3'-or 4',5'-positions and by following the metabolism of different radiolabeled inositol phosphates in extracts of RBL-2H3 cells. We report here that inositol 1,3,4,5-tetrakisphosphate, when incubated with the membrane fraction of extracts of RBL-2H3 cells, was converted to inositol 1,4,5-trisphosphate and inositol 1,3,4-trisphosphate. Further dephosphorylation of the inositol polyphosphates proceeded rapidly in whole extracts of cells, although the process was significantly retarded when ATP (2 mM) levels were maintained by an ATP-regenerating system. The degradation of inositol 1,4,5-trisphosphate proceeded with the sequential formation of inositol 1,4-bisphosphate, the inositol 4-monophosphate (with smaller amounts of the 1-monophosphate), and finally inositol. Inositol 1,3,4-trisphosphate, on the other hand, was converted to inositol 1,3-bisphosphate and inositol 3,4-bisphosphate and subsequently to inositol 4-monophosphate and inositol 1-monophosphate (stereoisomeric forms were undetermined). The possible implications of the apparent interconversion between inositol 1,4,5-trisphosphate and inositol 1,3,4,5-tetrakisphosphate in regulating histamine secretion in the RBL-2H3 cells are discussed.  相似文献   

10.
Preimplantation rabbit embryos collected at the early morula stage were cultured to blastocysts in the presence of [3H]inositol. The blastocysts were lysed, and both the aqueous and lipid portions were analysed for incorporated radioactivity. Thin-layer chromatographic separation of the lipid portion indicated that [3H]inositol was incorporated into phosphatidylinositol, phosphatidylinositol 4-phosphate, and phosphatidylinositol 4,5-bisphosphate. HPLC anion-exchange chromatography indicated that [3H]inositol was incorporated into inositol phosphates, including the two second messengers, inositol 1,4,5-trisphosphate and inositol 1,3,4,5-tetrakisphosphate, and also inositol monophosphate and inositol 1,4-bisphosphate. These results provide evidence that rabbit blastocysts may have an active phosphatidylinositol second messenger system, which may be responsive to intrauterine factors or intraembryonic paracrine factors. © 1993 Wiley-Liss, Inc.  相似文献   

11.
The metabolism of [3H]inositol (1,4,5)-trisphosphate was followed in permeabilized bovine adrenal glomerulosa cells. At low Ca++ concentration (pCa = 7.2), more than 90% of [3H]inositol (1,4,5)-trisphosphate had disappeared within 2 min, while two other metabolites, [3H]inositol (1,3,4)-trisphosphate and [3H]inositol (1,3,4,5)-tetrakisphosphate appeared progressively. At higher Ca++ concentrations (pCa = 5.7 and 4.8), the formation of these two metabolites was markedly increased, but completely abolished if the medium was ATP-depleted. The peak levels for the generation of [3H]inositol (1,3,4,5)-tetrakisphosphate (1 min) preceded those of [3H]inositol (1,3,4)-trisphosphate and were closely correlated. These results suggest that, in adrenal glomerulosa cells, the isomer inositol (1,3,4)-trisphosphate is generated from inositol (1,4,5)-trisphosphate via a calcium-sensitive and ATP-dependent phosphorylation/dephosphorylation pathway involving the formation of inositol (1,3,4,5)-tetrakisphosphate.  相似文献   

12.
In saponin-permeabilized mouse lacrimal acinar cells, glycerophosphoryl-myo-inositol 4,5-bisphosphate (GPIP2) activated the release of sequestered Ca2+ to the same extent as inositol 1,4,5-trisphosphate ((1,4,5)IP3) but with a potency about 1/10 that of (1,4,5)IP3. In lacrimal gland homogenates, [3H]GPIP2 was metabolized to two compounds which upon anion exchange high performance liquid chromatography eluted at positions indicating that they were [3H]GPIP and [3H]GPIP3. The rate of metabolism of [3H]GPIP2 was much slower than that of [3H](1,4,5)IP3, and its rate of phosphorylation was less than 1% of that of [3H] (1,4,5)IP3. In intact lacrimal cells, photolysis of a microinjected "caged" derivative of GPIP2, 1-(alpha-glycerophosphoryl)-myo-inositol 4,5-bisphosphate P4(5)-1-(2-nitrophenyl)ethyl ester, resulted in sustained activation of Ca2+ signaling; i.e. intracellular Ca2+ release followed by increased entry of Ca2+ across the plasma membrane. These findings indicate that caged GPIP2 should provide a useful tool for producing photolytically initiated, sustained activation of intracellular (1,4,5)IP3 receptors. They also provide strong support for the idea that sustained Ca2+ signaling can be achieved in lacrimal acinar cells by activation of intracellular receptors for (1,4,5)IP3 in the absence of stimulated production of inositol 1,3,4,5-tetrakisphosphate.  相似文献   

13.
We have synthesized two photolabile arylazido-analogues of Ins(1,4,5)P3 selectively substituted at the 1-phosphate group for determination of Ins(1,4,5)P3-binding proteins. These two photoaffinity derivatives, namely N-(4-azidobenzoyl)aminoethanol-1-phospho-D-myo-inositol 4,5-bisphosphate (AbaIP3) and N-(4-azidosalicyl)aminoethanol-1-phospho-D-myo-inositol 4,5-bisphosphate (AsaIP3), bind to high affinity Ins(1,4,5)P3-specific binding sites at a 9-fold lower affinity (Kd = 66 and 70 nM) than Ins(1,4,5)P3 (Kd = 7.15 nM) in a fraction from rat pancreatic acinar cells enriched in endoplasmic reticulum (ER). Other inositol phosphates tested showed comparable (DL-myo-inositol 1,4,5-trisphosphothioate, Kd = 81 nM) or much lower affinities for the binding sites [Ins(1,3,4,5)P4, Kd = 4 microM; Ins(1,4)P2, Kd = 80 microM]. Binding of AbaIP3 was also tested on a microsomal preparation of rat cerebellum [Kd = 300 nM as compared with Ins(1,4,5)P3, Kd = 45 nM]. Ca2+ release activity of the inositol derivatives was tested with AbaIP3. It induced a rapid and concentration-dependent Ca2+ release from the ER fraction [EC50 (dose producing half-maximal effect) = 3.1 microM] being only 10-fold less potent than Ins(1,4,5)P3 (EC50 = 0.3 microM). From the two radioactive labelled analogues ([3H]AbaIP3 and 125I-AsIP3) synthesized, the radioiodinated derivative was used for photoaffinity labelling. It specifically labelled three proteins with apparent molecular masses of 49, 37 and 31 kDa in the ER-enriched fraction. By subfractionation of this ER-enriched fraction on a Percoll gradient the 37 kDa Ins(1,4,5)P3 binding protein was obtained in a membrane fraction which showed the highest effect in Ins(1,4,5)P3-inducible Ca2+ release (fraction P1). The other two Ins(1,4,5)P3-binding proteins, of 49 and 31 kDa, were obtained in fraction P2, in which Ins(1,4,5)P3-induced Ca2+ release was half of that obtained in fraction P1. We conclude from these data that the 37 kDa and/or the 49 and 31 kDa proteins are involved in Ins(1,4,5)P3-induced Ca2+ release from the ER of rat pancreatic acinar cells.  相似文献   

14.
The action of carbachol on the generation of inositol trisphosphate and tetrakisphosphate isomers was investigated in dog-thyroid primary cultured cells radiolabelled with [3H]inositol. The separation of the inositol phosphate isomers was performed by reverse-phase high pressure liquid chromatography. The structure of inositol phosphates co-eluting with inositol 1,4,5-trisphosphate [Ins(1,4,5)P3] and inositol 1,3,4,5-tetrakisphosphate [Ins(1,3,4,5)P4] standards was determined by enzymatic degradation using a purified Ins(1,4,5)P3/Ins(1,3,4,5)P4 5-phosphatase. The data indicate that Ins(1,3,4,5)P4 was the only [3H]inositol phosphate which co-eluted with a [32P]Ins(1,3,4,5)P4 standard, whereas 80% of the [3H]InsP3 co-eluting with an Ins(1,4,5)P3 standard was actually this isomer. In the presence of Li+, carbachol led to rapid increases in [3H]Ins(1,4,5)P4. The level of Ins(1,4,5)P3 reached a peak at 200% of the control after 5-10 s of stimulation and fell to a plateau that remained slightly elevated for 2 min. The level of Ins(1,3,4,5)P4 reached its maximum at 20s. The level of inositol 1,3,4-trisphosphate [Ins(1,3,4)P3] increased continuously for 2 min after the addition of carbachol. Inositol-phosphate generation was also investigated under different pharmacological conditions. Li+ largely increased the level of Ins(1,3,4)P3 but had no effect on Ins(1,4,5)P3 and Ins(1,3,4,5)P4. Forskolin, which stimulates dog-thyroid adenylate cyclase and cyclic-AMP accumulation, had no effect on the generation of inositol phosphates. The absence of extracellular Ca2+ largely decreased the level of Ins(1,3,4,5)P4 as expected considering the Ca2(+)-calmodulin sensitivity of the Ins(1,4,5)P3 3-kinase. Staurosporine, an inhibitor of protein kinase C, increased the levels of Ins(1,4,5)P3, Ins(1,3,4,5)P4 and Ins(1,3,4)P3. This supports a negative feedback control of diacyglycerol on Ins(1,4,5)P3 generation.  相似文献   

15.
The kinetics of [3H]inositol phosphate metabolism in agonist-activated rat parotid acinar cells were characterized in order to determine the sources of [3H]inositol monophosphates and [3H]inositol bisphosphates. The turnover rates of D-myo-inositol 1,4,5-trisphosphate and its metabolites, D-myo-inositol 1,4-bisphosphate and D-myo-inositol 1,3,4-trisphosphate, were examined following the addition of the muscarinic receptor antagonist, atropine, to cholinergically stimulated parotid cells. D-myo-Inositol 1,4,5-trisphosphate declined with a t1/2 of 7.6 +/- 0.7 s, D-myo-inositol 1,3,4-trisphosphate declined with a t1/2 of 8.6 +/- 1.2 min, and D-myo-inositol 1,4-bisphosphate was metabolized with a t1/2 of 6.0 +/- 0.7 min. The sum of the rates of flux through D-myo-inositol 1,4-bisphosphate and D-myo-inositol 1,3,4-trisphosphate (2.54% phosphatidylinositol/min) did not exceed the calculated rate of breakdown of D-myo-inositol 1,4,5-trisphosphate (2.76% phosphatidylinositol/min). Thus, there is no evidence for the direct hydrolysis of phosphatidylinositol 4-phosphate in intact cells since D-myo-inositol 1,4-bisphosphate formation can be attributed to the dephosphorylation of D-myo-inositol 1,4,5-trisphosphate. The source of the [3H]inositol monophosphates also was examined in cholinergically stimulated parotid cells. When parotid cells were stimulated with methacholine, D-myo-inositol 1,4,5-trisphosphate, D-myo-inositol 1,3,4,5-tetrakisphosphate, D-myo-inositol 1,4-bisphosphate, and D-myo-inositol 4-monophosphate levels increased within 2 s, whereas D-myo-inositol 1-monophosphate accumulation was delayed by several seconds. Rates of [3H]inositol monophosphate accumulation also were examined by the addition of LiCl to cells stimulated to steady state levels of [3H]inositol phosphates. The sum of the rates of accumulation of D-myo-inositol 1-monophosphate and D-myo-inositol 4-monophosphate did not exceed the rate of breakdown of D-myo-inositol 1,4,5-trisphosphate or the sum of the rates of flux through D-myo-inositol 1,4-bisphosphate and D-myo-inositol 1,3,4-trisphosphate. These kinetic analyses suggest that agonist-stimulated [3H]inositol bis- and monophosphate formation in intact rat parotid acinar cells can be accounted for by the metabolism of D-myo-[3H]inositol 1,4,5-trisphosphate rather than by phospholipase C-catalyzed hydrolysis of phosphatidylinositol or phosphatidylinositol 4-phosphate.  相似文献   

16.
The hydrolysis of phosphatidylinositol may generate multiple second messengers, including inositol phosphates, 1,2-diacylglycerol, arachidonic acid, and phosphatidic acid. Here, we describe for the first time in spermatozoa that accumulation of one of these potential second messengers, inositol 1,4,5-trisphosphate (1,4,5-IP3), can be stimulated by the fucose-sulfate glycoconjugate (FSG) that induces an acrosome reaction. Sea urchin spermatozoa were labeled with myo-[3H]inositol and incubated with FSG. The amount of [3H]1,4,5-IP3 obtained from FSG-treated cells was up to 10 times that from untreated cells. Increases in the amount of [3H]1,4,5-IP3 were detected within 30 s after addition of FSG (2.5-fold) and were highest at 2 min after addition. Previously, it was shown that FSG induces Ca2+-dependent increases in cyclic AMP concentrations (Kopf, G. S., and Garbers, D. L. (1980) Biol. Reprod. 22, 1118-1126). Increases in [3H]1,4,5-IP3 accumulation caused by FSG were also dependent on extracellular Ca2+. The Ca2+ channel blockers, verapamil and nifedipine, inhibited increases in both [3H]1,4,5-IP3 and cyclic AMP, and the addition of concentrations of extracellular Ca2+ higher than 9.6 mM could reduce the inhibition. When spermatozoa were incubated in Ca2+-free seawater, FSG-induced increases in [3H]1,4,5-IP3 and cyclic AMP concentrations were blocked; addition of extracellular Ca2+ restored the responses. Other treatments that result in the induction of an acrosome reaction, including the addition of monovalent cation H+ exchangers, nigericin and gramicidin S, and incubation in seawater at alkaline pH (pH 8.8), also stimulated accumulation of [3H]1,4,5-IP3 and cyclic AMP.  相似文献   

17.
Inositol 1,3,4,5-tetrakisphosphates (Ins(1,3,4,5)P4), 32P-labelled in positions 4 and 5 were prepared enzymatically, using [4-32P]-phosphatidylinositol 4-phosphate (PtdInsP) and [5-32P]phosphatidylinositol 4,5-bisphosphate (PtdInsP2) as substrates, respectively. Degradation studies of Ins(1,3,4,5)P4, using an enriched phosphatase preparation from porcine brain cytosol, led to the formation of two inositol trisphosphate isomers which were identified as inositol 1,3,4-trisphosphate (Ins(1,3,4)P3) and inositol 1,4,5-trisphosphate (Ins(1,4,5)P3). This novel degradation pathway of Ins(1,3,4,5)P4 to Ins(1,4,5)P3 provides an additional source for the generation of Ins(1,4,5)P3, involving a 3-phosphatase.  相似文献   

18.
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   

19.
Rat hippocampal formation slices were prelabelled with [3H]inositol and stimulated with carbachol for times between 7 s and 3 min. The [3H]inositol metabolites in an acid extract of the slices were resolved with anion-exchange HPLC. Carbachol dramatically increased the concentration of [3H]inositol monophosphate, [3H]inositol bisphosphate (two isomers), [3H]inositol 1,3,4-trisphosphate, [3H]inositol 1,4,5-trisphosphate, and [3H]inositol 1,3,4,5-tetrakisphosphate. The levels of [3H]inositol 1,4,5-trisphosphate rose most rapidly; they were maximally elevated after only 7 s and declined toward control levels in 1 min followed by a more sustained elevation in levels for up to 3 min. When [3H]inositol 1,4,5-trisphosphate was incubated with hippocampal formation homogenates in an ATP-containing buffer it was very rapidly metabolised. After 5 min [3H]inositol 1,4-bisphosphate, [3H]inositol 1,3,4-trisphosphate, and [3H]inositol 1,3,4,5-tetrakisphosphate could be detected in the homogenates. Under similar experimental conditions [3H]inositol 1,3,4,5-tetrakisphosphate is metabolised to [3H]inositol 1,3,4-trisphosphate and an inositol bisphosphate isomer that is not [3H]inositol 1,4-bisphosphate. We conclude that like other tissues the primary event in the hippocampus following carbachol stimulation is the activation of phosphatidylinositol 4,5-bisphosphate selective phospholipase C.  相似文献   

20.
Binding of chemoattractants to specific cell surface receptors on polymorphonuclear leukocytes (PMNs) initiates a series of biochemical responses leading to cellular activation. A critical early biochemical event in chemoattractant (CTX) receptor-mediated signal transduction is the phosphodiesteric cleavage of plasma membrane phosphatidylinositol 4,5-bisphosphate (PIP2), with concomitant production of the calcium mobilizing inositol-1,4,5-trisphosphate (IP3) isomer, and the protein kinase C activator, 1,2-diacylglycerol (DAG). The following lines of experimental evidence collectively suggest that CTX receptors are coupled to phospholipase C via a guanine nucleotide binding (G) protein. Receptor-mediated hydrolysis of PIP2 in PMN plasma membrane preparations requires both fMet-Leu-Phe and GTP, and incubation of intact PMNs with pertussis toxin (which ADP ribosylates and inactivates some G proteins) eliminates the ability of fMet-Leu-Phe plus GTP to promote PIP2 breakdown in isolated plasma membranes. Studies with both PMN particulate fractions and with partially purified fMet-Leu-Phe receptor preparations indicate that guanine nucleotides regulate CTX receptor affinity. Finally, fMet-Leu-Phe stimulates high-affinity binding of GTP gamma S to PMN membranes as well as GTPase activity. A G alpha subunit has been identified in phagocyte membranes which is different from other G alpha subunits on the basis of molecular weight and differential sensitivity to ribosylation by bacterial toxins. Thus, a novel G protein may be involved in coupling CTX receptors to phospholipase C. Studies in intact and sonicated PMNs demonstrate that metabolism of 1,4,5-IP3 proceeds via two distinct pathways: 1) sequential dephosphorylation to 1,4-IP2, 4-IP1 and inositol, or 2) ATP-dependent conversion to inositol 1,3,4,5-tetrakisphosphate (IP4) followed by sequential dephosphorylation to 1,3,4-IP3, 3,4-IP2, 3-IP1 and inositol. Receptor-mediated hydrolysis of PIP2 occurs at ambient intracellular Ca2+ levels; but metabolism of 1,4,5-IP3 via the IP4 pathway requires elevated cytosolic Ca2+ levels associated with cellular activation. Thus, the two pathways for 1,4,5-IP3 metabolism may serve different metabolic functions. Additionally, inositol phosphate production appears to be controlled by protein kinase C, as phorbol myristate acetate (PMA) abrogates PIP2 hydrolysis by interfering with the ability of the activated G protein to stimulate phospholipase C. This implies a physiologic mechanism for terminating biologic responses via protein kinase C mediated feedback inhibition of PIP2 hydrolysis.  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