首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 62 毫秒
1.
Ca2+-activated Cl- channels are inhibited by inositol 3,4,5, 6-tetrakisphosphate (Ins(3,4,5,6)P4) (Xie, W., Kaetzel, M. A., Bruzik, K. S., Dedman, J. R., Shears, S. B., and Nelson, D. J. (1996) J. Biol. Chem. 271, 14092-14097), a novel second messenger that is formed after stimulus-dependent activation of phospholipase C (PLC). In this study, we show that inositol 1,3,4-trisphosphate (Ins(1,3,4)P3) is the specific signal that ties increased cellular levels of Ins(3,4,5,6)P4 to changes in PLC activity. We first demonstrated that Ins(1,3,4)P3 inhibited Ins(3,4,5,6)P4 1-kinase activity that was either (i) in lysates of AR4-2J pancreatoma cells or (ii) purified 22,500-fold (yield = 13%) from bovine aorta. Next, we incubated [3H]inositol-labeled AR4-2J cells with cell permeant and non-radiolabeled 2,5,6-tri-O-butyryl-myo-inositol 1,3, 4-trisphosphate-hexakis(acetoxymethyl) ester. This treatment increased cellular levels of Ins(1,3,4)P3 2.7-fold, while [3H]Ins(3, 4,5,6)P4 levels increased 2-fold; there were no changes to levels of other 3H-labeled inositol phosphates. This experiment provides the first direct evidence that levels of Ins(3,4,5,6)P4 are regulated by Ins(1,3,4)P3 in vivo, independently of Ins(1,3,4)P3 being metabolized to Ins(3,4,5,6)P4. In addition, we found that the Ins(1, 3,4)P3 metabolites, namely Ins(1,3)P2 and Ins(3,4)P2, were >100-fold weaker inhibitors of the 1-kinase compared with Ins(1,3,4)P3 itself (IC50 = 0.17 microM). This result shows that dephosphorylation of Ins(1,3,4)P3 in vivo is an efficient mechanism to "switch-off" the cellular regulation of Ins(3,4,5,6)P4 levels that comes from Ins(1,3, 4)P3-mediated inhibition of the 1-kinase. We also found that Ins(1,3, 6)P3 and Ins(1,4,6)P3 were poor inhibitors of the 1-kinase (IC50 = 17 and >30 microM, respectively). The non-physiological trisphosphates, D/L-Ins(1,2,4)P3, inhibited 1-kinase relatively potently (IC50 = 0.7 microM), thereby suggesting a new strategy for the rational design of therapeutically useful kinase inhibitors. Overall, our data provide new information to support the idea that Ins(1,3,4)P3 acts in an important signaling cascade.  相似文献   

2.
Ins(1,4,5,6)P4, a biologically active cell constituent, was recently advocated as a substrate of human Ins(3,4,5,6)P4 1-kinase (hITPK1), because stereochemical factors were believed relatively unimportant to specificity [Miller, G.J., Wilson, M.P., Majerus, P.W. and Hurley, J.H. (2005) Specificity determinants in inositol polyphosphate synthesis: crystal structure of inositol 1,3,4-triphosphate 5/6-kinase. Mol. Cell. 18, 201-212]. Contrarily, we provide three examples of hITPK1 stereospecificity. hITPK1 phosphorylates only the 1-hydroxyl of both Ins(3,5,6)P3 and the meso-compound, Ins(4,5,6)P3. Moreover, hITPK1 has >13,000-fold preference for Ins(3,4,5,6)P4 over its enantiomer, Ins(1,4,5,6)P4. The biological significance of hITPK1 being stereospecific, and not physiologically phosphorylating Ins(1,4,5,6)P4, is reinforced by our demonstrating that Ins(1,4,5,6)P4 is phosphorylated (K(m) = 0.18 microM) by inositolphosphate-multikinase.  相似文献   

3.
Regulation of Cl(-) channel conductance by Ins(3,4,5,6)P(4) provides receptor-dependent control over salt and fluid secretion, cell volume homeostasis, and electrical excitability of neurones and smooth muscle. Ignorance of how Ins(3,4,5,6)P(4) is synthesized has long hindered our understanding of this signaling pathway. We now show Ins(3,4,5,6)P(4) synthesis by Ins(1,3,4,5,6)P(5) 1-phosphatase activity by an enzyme previously characterized as an Ins(3,4,5,6)P(4) 1-kinase. Rationalization of these phenomena with a ligand binding model unveils Ins(1,3,4)P(3) as not simply an alternative kinase substrate, but also an activator of Ins(1,3,4,5,6)P(5) 1-phosphatase. Stable overexpression of the enzyme in epithelial monolayers verifies its physiological role in elevating Ins(3,4,5,6)P(4) levels and inhibiting secretion. It is exceptional for a single enzyme to catalyze two opposing signaling reactions (1-kinase/1-phosphatase) under physiological conditions. Reciprocal coordination of these opposing reactions offers an alternative to general doctrine that intracellular signals are regulated by integrating multiple, distinct phosphatases and kinases.  相似文献   

4.
Avian erythrocytes were incubated with myo-[3H]inositol for 6-7 h and with [32P]Pi for the final 50-90 min of this period. An acid extract was prepared from the prelabelled erythrocytes, and the specific radioactivities of the gamma-phosphate of ATP and of both the myo-inositol moieties (3H, d.p.m./nmol) and the individual phosphate groups (32P, d.p.m./nmol) of [3H]Ins[32P](1,3,4,6)P4,[3H]Ins[32P](1,3,4,5)P4, [3H]Ins[32P](3,4,5,6)P4 and [3H]Ins[32P](1,3,4,5,6)P5 were determined. The results provide direct confirmation that one of the cellular InsP4 isomers is Ins(1,3,4,5)P4 which is synthesized by sequential phosphorylation of the 1,4,5 and 3 substitution sites of the myo-Ins moiety, precisely as previously deduced [Batty, Nahorski & Irvine (1985) Biochem. J. 232, 211-215; Irvine, Letcher, Heslop & Berridge (1986) Nature (London) 320, 631-634]. This is compatible with the proposed synthetic route from PtdIns via PtdIns4P, PtdIns(4,5)P2 and Ins(1,4,5)P3. The data also suggest that, in avian erythrocytes, the principle precursor of Ins(1,3,4,5,6)P5 is Ins(3,4,5,6)P4. Furthermore, if the gamma- (and/or beta-) phosphate of ATP is the precursor of the phosphate moieties of Ins(3,4,5,6)P4, then this isomer must be derived from the phosphorylation of Ins(3,4,6)P3. If the gamma- (and/or beta-) phosphate of ATP similarly acts as the ultimate precursor to all of the phosphates of Ins(1,3,4,6)P4, then, in intact avian erythrocytes, the main precursor of Ins(1,3,4,6)P4 is Ins(1,4,6)P3. This contrasts with the expectation, based on results with cell-free systems, that Ins(1,3,4,6)P4 is synthesized by the direct phosphorylation of Ins(1,3,4)P3.  相似文献   

5.
Does inositol 3,4,5,6-tetrakisphosphate (Ins(3,4,5,6)P(4)) inhibit apical Ca(2+)-activated Cl(-) conductance (CaCC)? We studied this question using human CFPAC-1 pancreatoma cells grown in polarized monolayers. Cellular Ins(3,4,5,6)P(4) levels were acutely sensitive to purinergic receptor activation, rising 3-fold within 1 min of agonist addition. Intracellular Ins(3,4,5,6)P(4) levels were therefore specifically elevated, independently of receptor activation, by incubating cells with a cell-permeant bioactivable analogue, 1,2-di-O-butyl-myo-inositol 3,4,5,6-tetrakisphosphate octakis(acetoxymethyl)ester (Bt(2)Ins (3,4,5,6)P(4)/AM). The latter inhibited Ca(2+)-activated Cl(-) secretion by 60%. We next used nystatin to selectively permeabilize the basolateral membrane to monovalent anions and cations, thereby preventing this membrane from electrochemically dominating ion movements through the apical membrane. Thus, we studied autonomous regulation of apical Cl(-) channels in situ. The properties of Cl(-) flux across the apical membrane were those expected of CaCC: niflumic acid sensitivity, outward rectification, and 2-fold greater permeability of I(-) over Cl(-). Following nystatin-treatment, we elevated intracellular levels of Ins(3,4,5,6)P(4) with either purinergic agonists or with Bt(2)Ins(3,4,5,6)P(4)/AM. Both protocols inhibited Ca(2+)-activated Cl(-) secretion (up to 70%). These studies provide the first demonstration that, in a physiologically relevant context of a polarized monolayer, there is an apical, Ins(3,4,5,6)P(4)-inhibited CaCC.  相似文献   

6.
Cystic fibrosis (CF) patients suffer from a defect in hydration of mucosal membranes due to mutations in the cystic fibrosis transmembrane regulator (CFTR), an apical chloride channel in mucosal epithelia. Disease expression in CF knockout mice is organ specific, varying with the level of expression of calcium activated Cl(-) channels (CLCA). Therefore, restoring transepithelial Cl(-) secretion by augmenting alternate Cl(-) channels, such as CLCA, could be beneficial. However, CLCA-mediated Cl(-) secretion is transient, due in part to the inhibitory effects of myo-inositol 3,4,5,6-tetrakisphosphate [Ins(3,4,5,6)P(4)]. This suggests that antagonists of Ins(3,4,5,6)P(4) could be useful in treatment of CF. We have, therefore, synthesized a series of membrane-permeant Ins(3,4,5,6)P(4) derivatives, carrying alkyl substituents on the hydroxyl groups and screened them for effects on Cl(-) secretion in a human colonic epithelial cell line, T(84). While membrane-permeant Ins(3,4,5,6)P(4) derivatives had no direct effects on carbachol-stimulated Cl(-) secretion, Ins(3,4,5,6)P(4) derivatives, but not enantiomeric Ins(1,4,5,6)P(4) derivatives, reversed the inhibitory effect of Ins(3,4,5,6)P(4) on subsequent thapsigargin activation of Cl(-) secretion. The extent of the antagonistic effect of the Ins(3,4,5,6)P(4) derivatives varied with the position of the alkyl substituents. Derivatives with a cyclohexylidene ketal or a butyl-chain at the 1-position reversed the Ins(3,4,5,6)P(4)-mediated inhibition of Cl(-) secretion by up to 96 and 85%, respectively, whereas butylation of the 1- and 2-position generated a reversal effect of only 65%. Derivatives carrying the butyl chain only at the 2-position showed no antagonistic effect. These data: (1) Support the hypothesis that Ins(3,4,5,6)P(4) stereospecifically inhibits Ca(2+) activated Cl(-) secretion and that Ins(3,4,5,6)P(4) mediates most, if not all of the cholinergic-mediated inhibition of chloride secretion in T(84) cells; (2) Demonstrate Ins(3,4,5,6)P(4)-mediated inhibition can be completely reversed with rationally designed membrane-permeant Ins(3,4,5,6)P(4) antagonists; (3) Demonstrate that a SAR for membrane-permeant Ins(3,4,5,6) P(4) antagonists can be generated and screened in a physiologically relevant cell-based assay; (4) Indicate that Ins(3,4,5,6)P(4) derivatives could serve as a starting point for the development of therapeutics to treat cystic fibrosis.  相似文献   

7.
Oscillatory growth of pollen tubes has been correlated with oscillatory influxes of the cations Ca(2+), H(+), and K(+). Using an ion-specific vibrating probe, a new circuit was identified that involves oscillatory efflux of the anion Cl(-) at the apex and steady influx along the tube starting at 12 microm distal to the tip. This spatial coupling of influx and efflux sites predicts that a vectorial flux of Cl(-) ion traverses the apical region. The Cl(-) channel blockers 4,4'-diisothiocyanatostilbene-2,2'-disulfonic acid (DIDS) and 5-nitro-2-(3-phenylpropylamino)benzoic acid completely inhibited tobacco pollen tube growth at 80 and 20 microM, respectively. Cl(-) channel blockers also induced increases in apical cell volume. The apical 50 micro m of untreated pollen tubes had a mean cell volume of 3905 +/- 75 microm(3). DIDS at 80 microM caused a rapid and lethal cell volume increase to 6206 +/- 171 microm(3), which is at the point of cell bursting at the apex. DIDS was further demonstrated to disrupt Cl(-) efflux from the apex, indicating that Cl(-) flux correlates with pollen tube growth and cell volume status. The signal encoded by inositol 3,4,5,6-tetrakisphosphate [Ins(3,4,5,6)P(4)] antagonized pollen tube growth, induced cell volume increases, and disrupted Cl(-) efflux. Ins(3,4,5,6)P(4) decreased the mean growth rate by 85%, increased the cell volume to 5997 +/- 148 microm(3), and disrupted normal Cl(-) efflux oscillations. These effects were specific for Ins(3,4,5,6)P(4) and were not mimicked by either Ins(1,3,4,5)P(4) or Ins(1,3,4,5,6)P(5). Growth correlation analysis demonstrated that cycles of Cl(-) efflux were coupled to and temporally in phase with cycles of growth. A role for Cl(-) flux in the dynamic cellular events during growth is assessed. Differential interference contrast microscopy and kymographic analysis of individual growth cycles revealed that vesicles can advance transiently to within 2 to 4 microm of the apex during the phase of maximally increasing Cl(-) efflux, which temporally overlaps the phase of cell elongation during the growth cycle. In summary, these investigations indicate that Cl(-) ion dynamics are an important component in the network of events that regulate pollen tube homeostasis and growth.  相似文献   

8.
Ins(3,4,5,6)P(4) inhibits plasma membrane Cl(-) flux in secretory epithelia [1]. However, in most other mammalian cells, receptor-dependent elevation of Ins(3,4,5,6)P(4) levels is an "orphan" response that lacks biological significance [2]. We set out to identify Cl(-) channel(s) and/or transporter(s) that are regulated by Ins(3,4,5,6)P4 in vivo. Several candidates [3-5] were excluded through biophysical criteria, electrophysiological analysis, and confocal immunofluorescence microscopy. Then, we heterologously expressed ClC-3 in the plasma membrane of HEK293-tsA201 cells; whole-cell patch-clamp analysis showed Ins(3,4,5,6)P4 to inhibit Cl(-) conductance through ClC-3. Next, we heterologously expressed ClC-3 in the early endosomal compartment of BHK cells; by fluorescence ratio imaging of endocytosed FITC-transferrin, we recorded intra-endosomal pH, an in situ biosensor for Cl(-) flux across endosomal membranes [6]. A cell-permeant, bioactivatable Ins(3,4,5,6)P4 analog elevated endosomal pH from 6.1 to 6.6, reflecting inhibition of ClC-3. Finally, Ins(3,4,5,6)P(4) inhibited endogenous ClC-3 conductance in postsynaptic membranes of neonatal hippocampal neurones. Among other ClC-3 functions that could be regulated by Ins(3,4,5,6)P4 are tumor cell migration [7], apoptosis [8], and inflammatory responses [9]. Ins(3,4,5,6)P4 is a ubiquitous cellular signal with diverse biological actions.  相似文献   

9.
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   

10.
We have studied the regulation of Ca(2+)-dependent chloride (Cl(Ca)) channels in a human pancreatoma epithelial cell line (CFPAC-1), which does not express functional cAMP-dependent cystic fibrosis transmembrane conductance regulator chloride channels. In cell-free patches from these cells, physiological Ca(2+) concentrations activated a single class of 1-picosiemens Cl(-)-selective channels. The same channels were also stimulated by a purified type II calmodulin-dependent protein kinase (CaMKII), and in cell-attached patches by purinergic agonists. In whole-cell recordings, both Ca(2+)- and CaMKII-dependent mechanisms contributed to chloride channel stimulation by Ca(2+), but the CaMKII-dependent pathway was selectively inhibited by inositol 3,4,5,6-tetrakisphosphate (Ins(3,4,5,6)P(4)). This inhibitory effect of Ins(3,4,5,6)P(4) on Cl(Ca) channel stimulation by CaMKII was reduced by raising [Ca(2+)] and prevented by inhibition of protein phosphatase activity with 100 nm okadaic acid. These data provide a new context for understanding the physiological relevance of Ins(3,4,5,6)P(4) in the longer term regulation of Ca(2+)-dependent Cl(-) fluxes in epithelial cells.  相似文献   

11.
The metabolism of inositol 1,3,4-trisphosphate to inositol 1,3-bisphosphate   总被引:9,自引:0,他引:9  
We previously demonstrated a pathway for the metabolism of inositol 1,3,4-trisphosphate (Ins(1,3,4)P3) to inositol 3,4-bisphosphate (Ins(3,4)P2) in calf brain extracts. Inositol polyphosphate 1-phosphatase, a Mg2+-dependent, lithium ion-inhibited enzyme, specifically hydrolyzes Ins(1,3,4)P3 to Ins(3,4)P2 and Ins(1,4)P2 to Ins 4-P (Inhorn, R. C., Bansal, V. S., and Majerus, P. W. (1987) Proc. Natl. Acad. Sci. U.S.A. 84, 2170-2174). Now we have found an alternative pathway for the metabolism of Ins(1,3,4)P3 in crude calf brain extracts. Along this pathway, Ins(1,3,4)P3 is first converted to Ins(1,3)P2 which is further hydrolyzed to Ins 1-P. This pathway involves a 4-phosphatase and a 3-phosphatase which do not require Mg2+ and are not inhibited by lithium ions. A similar 4-phosphatase also degrades Ins(3,4)P2 to Ins 3-P. Three different inositol bisphosphates formed from calf brain supernatant are each further metabolized by a separate enzyme. The three inositol monophosphates, i.e. Ins 1-P, Ins 3-P, and Ins 4-P, are converted to inositol by inositol monophosphate phosphatase (Ackermann, K. E., Gish, B. G., Honchar, M. P., and Sherman, W. R. (1987) Biochem. J. 242, 517-524).  相似文献   

12.
Pasteurella multocida toxin, both native and recombinant, is an extremely potent mitogen for Swiss 3T3 cells and acts to enhance the formation of total inositol phosphates (Rozengurt, E., Higgins, T., Changer, N., Lax, A.J., and Staddon, J.M. (1990) Proc. Natl. Acad. Sci. U.S.A. 87, 123-127). P. multocida toxin also stimulates diacylglycerol production and activates protein kinase C (Staddon, J.M., Chanter, N., Lax, A.J., Higgins, T.E., and Rozengurt, E. (1990) J. Biol. Chem. 265, 11841-11848). Here we analyze, by [3H]inositol labeling and high performance liquid chromatography, the inositol phosphates in recombinant P. multocida toxin-treated cells. Recombinant P. multocida toxin stimulated increases in [3H]inositol 1,4,5-trisphosphate ([3H]Ins(1,4,5)P3) and its metabolic products, including Ins(1,3,4,5)P4, Ins(1,3,4)P3, Ins(1,4)P2, Ins(4/5)P, and Ins(1/3)P. The profile of the increase in the cellular content of these distinct inositol phosphates was very similar to that elicited by bombesin. Furthermore, recombinant P. multocida toxin, like bombesin, mobilizes an intracellular pool of Ca2+. Recombinant P. multocida toxin pretreatment greatly reduces the Ca2(+)-mobilizing action of bombesin, consistent with Ca2+ mobilization from a common pool by the two agents. The enhancement of inositol phosphates and mobilization of Ca2+ by recombinant P. multocida toxin were blocked by the lysosomotrophic agents methylamine, ammonium chloride, and chloroquine and occurred after a dose-dependent lag period. The stimulation of inositol phosphate production by recombinant P. multocida toxin persisted after removal of extracellular toxin, in contrast to the reversibility of the action of bombesin. Recombinant P. multocida toxin, unlike bombesin and guanosine 5'-O-(gamma-thiotriphosphate), did not cause the release of inositol phosphates in permeabilized cells. These data demonstrate that recombinant P. multocida toxin, acting intracellularly, stimulates the phospholipase C-mediated hydrolysis of phosphatidylinositol 4,5-bisphosphate.  相似文献   

13.
Ins(3,4,5,6)P(4) is an inhibitor of Ca(2+)-activated Cl(-) channels, but further understanding has been hindered by ignorance of how it is made in cells. It now transpires that one protein with ATP-dependent kinase and phosphatase activities interconverts Ins(3,4,5,6)P(4) and Ins(1,3,4,5,6)P(5), as well as several other inositol polyphosphates.  相似文献   

14.
Metabolic and signaling properties of an Itpk gene family in Glycine max   总被引:2,自引:0,他引:2  
Stiles AR  Qian X  Shears SB  Grabau EA 《FEBS letters》2008,582(13):1853-1858
We have cloned and characterized four Itpk genes from soybean. All four recombinant Itpk proteins showed canonical Ins(1,3,4)P3 5/6-kinase activity, but a kinetic analysis raised questions about its biological significance. Instead, we provide evidence that one alternative biological role for soybean Itpks is to interconvert the Cl(-) channel inhibitor, Ins(3,4,5,6)P4, and its metabolic precursor, Ins(1,3,4,5,6)P5, within a substrate cycle. The soybean Itpks also phosphorylated Ins(3,4,6)P3 to Ins(1,3,4,6)P4 which was further phosphorylated to Ins(1,3,4,5,6)P5 by soybean Ipk2. Thus, soybean Itpks may participate in an inositol lipid-independent pathway of InsP6 synthesis.  相似文献   

15.
C. reinhardtii sheds its flagella in response to acidification. Previously, we showed correlations between pH shock, deflagellation, and inositol 1,4,5-trisphosphate [Ins(1,4,5)P3] production, but 100% of cells deflagellated by 5 s, which was the earliest that Ins(1,4,5)P3 accumulation could be accurately measured by techniques available to us at that time (Quarmby, L. M., Y. G. Yueh, J. L. Cheshire, L. R. Keller, W. J. Snell, and R. C. Crain. J. Cell Biol. 1992. 116:737-744). To learn about the causal relationship between Ins(1,4,5)P3 accumulation and deflagellation, we extended these studies to early times using a continuous-flow rapid-quench device. Within 1 s of acidification to pH 4.3-4.5, 100% of cells deflagellated. A transient peak of Ins(1,4,5)P3 was observed 250-350 ms after pH shock, preceding deflagellation. Preincubation with 10 microM neomycin, which prevents hydrolysis of phosphatidylinositol 4,5-bisphosphate, inhibited both the transient production of Ins(1,4,5)P3 and the subsequent deflagellation. The nonspecific Ca2+ channel blockers La3+ and Cd2+ prevented flagellar excision induced by mastoparan without inhibiting rapid Ins(1,4,5)P3 production. Likewise, the Ins(1,4,5)P3-gated channel inhibitors ruthenium red and heparin blocked deflagellation in response to mastoparan. These studies were extended to mutants defective in flagellar excision. Fa-1, a mutant defective in flagellar structure, produced Ins(1,4,5)P3 but failed to deflagellate. These results support a model in which acid pH activates a putative cellular receptor leading to G-protein dependent activation of phospholipase C and accumulation of Ins(1,4,5)P3. These events are upstream of Ins(1,4,5)P3-dependent Ca2+ entry from the medium, and of deflagellation.  相似文献   

16.
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   

17.
Considerable attention has focused on the role of protein kinase C (PKC) in triggering the profound infarct-sparing effect of ischemic preconditioning (PC). In contrast, the involvement of inositol 1,4,5-trisphosphate [Ins(1,4,5)P(3)], the second messenger generated in parallel with the diacylglycerol-PKC pathway, remains poorly understood. We hypothesized that, if Ins(1,4,5)P(3) signaling [i.e., release of Ins(1,4,5)P(3) and subsequent binding to Ins(1,4,5)P(3) receptors] contributes to PC-induced cardioprotection, then the reduction of infarct size achieved with PC would be attenuated in mice that are deficient in Ins(1,4,5)P(3) receptor protein. To test this concept, hearts were harvested from 1) B6C3Fe-a/a-Itpr-1(opt+/-)/J mutants displaying reduced expression of Ins(1,4,5)P(3) receptor-1 protein, 2) Itpr-1(opt+/+) wild types from the colony, and 3) C57BL/6J mice. All hearts were buffer-perfused and randomized to receive two 5-min episodes of PC ischemia, pretreatment with d-myo-Ins(1,4,5)P(3) [sodium salt of native Ins(1,4,5)P(3)], the mitochondrial ATP-sensitive K(+) channel opener diazoxide, or no intervention (controls). After the treatment phase, all hearts underwent 30-min global ischemia followed by 2 h of reperfusion, and infarct size was delineated by tetrazolium staining. In both wild-type and C57BL/6J cohorts, area of necrosis in hearts that received PC, d-myo-Ins(1,4,5)P(3), and diazoxide averaged 28-35% of the total left ventricle (LV), significantly smaller than the values of 52-53% seen in controls (P < 0.05). In contrast, in Itpr-1(opt+/-) mutants, protection was only seen with diazoxide: neither PC nor d-myo-Ins(1,4,5)P(3) limited infarct size (52-58% vs. 56% of the LV in mutant controls). These data provide novel evidence that Ins(1,4,5)P(3) signaling contributes to infarct size reduction with PC.  相似文献   

18.
In previous works, we synthesized a series of inositol 1,4,5-trisphosphate (Ins(1,4,5)P3) analogs, with a substituent on the second carbon of the inositol ring. Using these analogs, the Ins(1,4,5)P3 affinity media were also synthesized (Hirata, M., Watanabe, Y., Ishimatsu, T., Yanaga, F., Koga, T., and Ozaki, S. (1990) Biochem. Biophys. Res. Commun. 168, 379-386). When the cytosol fraction from the rat brain was applied to an Ins(1,4,5)P3 affinity column, an eluate with a 2 M NaCl solution was found to have remarkable Ins(1,4,5)P3-binding activity. The active fraction was further fractionated with gel filtration chromatography, and two proteins with an apparent molecular mass of 130 or 85 kDa were found to be Ins(1,4,5)P3-binding proteins but with no Ins(1,4,5)P3 metabolizing activities. Partial amino acid sequences determined after proteolysis and reversed-phase chromatography revealed that the protein with an apparent molecular mass of 85 kDa is the delta-isozyme of phospholipase C and that of 130 kDa has no sequence the same as the Ins(1,4,5)P3-recognizing proteins hitherto examined. Ins(1,4,5)P3 at concentrations greater than 1 microM strongly inhibited 85-kDa phospholipase C delta activity, without changing its dependence on the concentrations of free Ca2+ and H+. Among inositol phosphates examined, Ins(3,4,5,6)P4 inhibited the binding of [3H]Ins(1,4,5)P3 to the 130-kDa protein at much the same concentrations as seen with Ins(1,4,5)P3. This report seems to be the first evidence for the presence of soluble Ins(1,4,5)P3-binding proteins in the rat brain, one of which is the delta isozyme of phospholipase C.  相似文献   

19.
PRIP-1 was isolated as a novel inositol 1,4,5-trisphosphate [Ins(1,4,5)P3] binding protein with a domain organization similar to phospholipase C-delta1 (PLC-delta1) but lacking the enzymatic activity. Further studies revealed that the pleckstrin homology (PH) domain of PRIP-1 is the region responsible for binding Ins(1,4,5)P3. In this study we aimed to clarify the role of PRIP-1 at the physiological concentration in Ins(1,4,5)P3-mediated Ca2+ signaling, as we had previously used COS-1 cells overexpressing PRIP-1 (Takeuchi et al., 2000, Biochem J 349:357-368). For this purpose we employed PRIP-1 knock out (PRIP-1-/-) mice generated previously (Kanematsu et al., 2002, EMBO J 21:1004-1011). The increase in free Ca2+ concentration in response to purinergic receptor stimulation was lower in primary cultured cortical neurons prepared from PRIP-1-/- mice than in those from wild type mice. The relative amounts of [3H]Ins(1,4,5)P3 measured in neurons labeled with [3H]inositol was also lower in cells from PRIP-1-/- mice. In contrast, PLC activities in brain cortex samples from PRIP-1-/- mice were not different from those in the wild type mice, indicating that the hydrolysis of Ins(1,4,5)P3 is enhanced in cells from PRIP-1-/- mice. In vitro analyses revealed that type1 inositol polyphosphate 5-phosphatase physically interacted with a PH domain of PRIP-1 (PRIP-1PH) and its enzyme activity was inhibited by PRIP-1PH. However, physical interaction with these two proteins did not appear to be the reason for the inhibition of enzyme activity, indicating that binding of Ins(1,4,5)P3 to the PH domain prevented its hydrolyzation. Together, these results indicate that PRIP-1 plays an important role in regulating the Ins(1,4,5)P3-mediated Ca2+ signaling by modulating type1 inositol polyphosphate 5-phosphatase activity through binding to Ins(1,4,5)P3.  相似文献   

20.
In a variety of cell types, insulin stimulation elicits the rapid production of H(2)O(2), which causes the oxidative inhibition of protein-tyrosine phosphatases and enhances the tyrosine phosphorylation of proteins in the early insulin action cascade (Mahadev, K., Zilbering, A., Zhu, L., and Goldstein, B. J. (2001) J. Biol. Chem. 276, 21938-21942). In the present work, we explored the potential role of insulin-induced H(2)O(2) generation on downstream insulin signaling using diphenyleneiodonium (DPI), an inhibitor of cellular NADPH oxidase that blocks insulin-stimulated cellular H(2)O(2) production. DPI completely inhibited the activation of phosphatidylinositol (PI) 3'-kinase activity by insulin and reduced the insulin-induced activation of the serine kinase Akt by up to 49%; these activities were restored when H(2)O(2) was added back to cells that had been pretreated with DPI. Interestingly, the H(2)O(2)-induced activation of Akt was entirely mediated by upstream stimulation of PI 3'-kinase activity, since treatment of 3T3-L1 adipocytes with the PI 3'-kinase inhibitors wortmannin or LY294002 completely blocked the subsequent activation of Akt by exogenous H(2)O(2). Preventing oxidant generation with DPI also blocked insulin-stimulated glucose uptake and GLUT4 translocation to the plasma membrane, providing further evidence for an oxidant signal in the regulation of the distal insulin-signaling cascade. Finally, in contrast to the cellular mechanism of H(2)O(2) generation by other growth factors, such as platelet-derived growth factor, we also found that insulin-stimulated cellular production of H(2)O(2) may occur through a unique pathway, independent of cellular PI 3'-kinase activity. Overall, these data provide insight into the physiological role of insulin-dependent H(2)O(2) generation, which is not only involved in the regulation of tyrosine phosphorylation events in the early insulin signaling cascade but also has important effects on the regulation of downstream insulin signaling, involving the activation of PI 3'-kinase, Akt, and ultimately cellular glucose transport in response to insulin.  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