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.
Summary The latent period before depolarization of Limulus ventral photoreceptors by light flashes was compared with that following brief, intracellular, pressure-injection of d-myo-inositol 1,4,5 trisphosphate. At temperatures between 18 °C and 22 °C and with an extracellular calcium concentration of 10 mM, the responses of 4 cells to light and to injections of 100 M inositol trisphosphate displayed average latencies of 71 and 56 ms, respectively. The latencies of responses to InsP3 included an estimated 20 ms dead-time inherent in the injection method. Reducing the temperature lengthened the latency of the response to light (Q10 approximately 3.2 between 7 and 22 °C) more than that to inositol trisphosphate (Q10 approximately 2.3). Bathing the photoreceptors in seawater containing no added calcium and 1 mM of the calcium chelator EGTA greatly increased the latency of the light response at all temperatures, but did not increase the latency of the response to inositol trisphosphate. We conclude that the response to inositol trisphosphate lacks the calcium- and temperature-sensitive latent period which characterizes the response to light. If inositol trisphosphate acts, via the release of stored calcium, to stimulate an intermediate in the visual cascade, then that intermediate would appear to be downstream from the latency-generating mechanism.Abbreviations InsP 3 D-myo-inositol 1,4,5 trisphosphate - ASW Artificial seawater - Ca i Cytosolic free calcium ion concentration - Ca 0 Extracellular calcium ion concentration  相似文献   

2.
Several studies have demonstrated the activation of phosphoinositide-specific phospholipase C (Plc) in nuclei of mammalian cells during synchronous progression through the cell cycle, but the downstream targets of Plc-generated inositol 1,4,5-trisphosphate are poorly described. Phospholipid signaling in the budding yeast Saccharomyces cerevisiae shares similarities with endonuclear phospholipid signaling in mammals, and many recent studies point to a role for inositol phosphates, including InsP5, InsP6, and inositol pyrophosphates, in mediating the action of Plc. In this study, we investigated the changes in inositol phosphate levels in α-factor-treated S. cerevisiae, which allows cells to progress synchronously through the cell cycle after release from a G1 block. We found an increase in the activity of Plc1 early after release from the block with a concomitant increase in the levels of InsP7 and InsP8. Treatment of cells with the Plc inhibitor U73122 prevented increases in inositol phosphate levels and blocked progression of cells through S phase after pheromone arrest. The enzymatic activity of Kcs1 in vitro and HPLC analysis of [3H]inositol-labeled kcs1Δ cells confirmed that Kcs1 is the principal kinase responsible for generation of pyrophosphates in synchronously progressing cells. Analysis of plc1Δ, kcs1Δ, and ddp1Δ yeast mutants further confirmed the role that a Plc1- and Kcs1-mediated increase in pyrophosphates may have in progression through S phase. Our data provide genetic, metabolic, and biochemical evidence that synthesis of inositol pyrophosphates through activation of Plc1 and Kcs1 plays an important role in the signaling response required for cell cycle progression after mating pheromone arrest.  相似文献   

3.
Abstract: The regional distribution of inositol 1,4,5-trisphosphate (InsP3), inositol 1,3,4,5-tetrakisphosphate (InsP4), and ryanodine binding sites has been characterised and compared in the rat brain using radioligand binding assays. Cortical [3H]InsP3 binding indicated similar positional and stereospecificity as observed in other tissues, with 100-fold selectivity for lnsP3 over InsP4. Similarly, high-affinity [32P]InsP4 binding also showed a high degree of positional specificity, with a 1,000-fold selectivity for InsP4 over InsP3. Initial characterisation of [3H]ryanodine binding to cortical membranes demonstrated that specific binding was highly dependent on high salt and micromolar Ca2+ concentrations and inhibited by Ca2+ levels of >1 mM. [3H]-Ryanodine binding was also enhanced by β,γ-methylene-adenosine 5′-trisphosphate and caffeine and inhibited by magnesium and ruthenium red (Ki= 0.81 μM). However, dantrolene (300 μM) was ineffective on the binding. Therefore, although the results indicate a greater similarity to the binding properties of the Ca2+-induced Ca2+ release channel isoform present in skeletal, rather than cardiac, muscle, it does not appear to be identical. Detailed binding analysis of ryanodine and polyphosphate sites, with the exception of ruthenium red, indicated no interaction between binding sites. Ruthenium red markedly enhanced the binding of both [3H]InsP3 and [32P]InsP4, an effect most probably due to nonspecific complex formation. Regional binding of InP3, InsP4, and ryanodine in the rat brain was of similar affinity for each ligand in each area, but the density profile for each ligand was clearly different. The highest density of InsP3 sites was in the cerebellum, whereas the highest density of ryanodine sites was in the hippocampus. High-affinity InsP4 sites showed less regional diversity, with highest densities in the cerebellum, cortex, and hippocampus. However, in each area studied the density of sites followed the order InsP3 > InsP4 > ryanodine.  相似文献   

4.
The rate of synthesis of inositol trisphosphate (InsP3) in trophocytes derived from disaggregated cockroach (Periplaneta americana) fat body increases following treatment of the cells with hypertrehalosemic hormone I or II (HTH-I, -II) in vitro. Trophocytes preloaded with [3H]inositol display a significant increase in InsP3 synthesis as early as 15 s after addition of the hormone. When the trophocytes are pre-incubated with LiCl and subsequently incubated with HTH the [3H] content of the InsP3 fraction is greater than that found with HTH alone. This is taken as evidence that inositol monophosphate phosphatase is part of the mechanism for clearing InsP3 from the cytosol. In contrast to HTH, octopamine, which is also capable of exerting a hypertrehalosemic effect in the cockroach, does not increase the synthesis of InsP3. 1-Octadecyl-2-methyl-rac-glycero-3-phosphocholine (ET-18-OCH3), a potent and selective inhibitor of phosphatidylinositol phospholipase C, blocks the activation of phosphorylase by HTH-I as well as the hypertrehalosemic effect induced by the hormone.  相似文献   

5.
Within the first 2 h of sexual reproduction, gametes of the green alga Chlamydomonas eugametos agglutinate, fuse via their mating structures, de-agglutinate and swim off as vis-à-vis pairs. During this period, increases in intracellular inositol 1,4,5-trisphosphate levels and changes in polyphosphoinositide synthesis were associated with cell fusion. The protein-kinase-C inhibitor staurosporine (0.1–0.2 M) inhibited the de-agglutination of pairs and therefore prevented them swimming away, while earlier stages of mating such as agglutination or cell fusion were unaffected. The results suggest that inositol 1,4,5-trisphosphate and diacylglycerol are fertilization signals in C. eugametos. The idea that they could also be fertilization signals in higher plants is discussed in relation to in vitro embryogenesis.Abbreviations DAG diacylglycerol - InsP3 inositol 1,4,5-trisphosphate - mt+/mt mating-type plus or minus - PKC protein kinase C - PtdOH phosphatidic acid - PtdInsP phosphatidylinositol - 4-phosphate PtdInsP2 phosphatidylinositol 4,5-bisphosphate  相似文献   

6.
InsP3-mediated calcium release through the type 2 inositol 1,4,5-trisphosphate receptor (InsP3R2) in cardiac myocytes results in the activation of associated CaMKII, thus enabling the kinase to act on downstream targets, such as histone deacetylases 4 and 5 (HDAC4 and HDAC5). The CaMKII activity also feedback modulates InsP3R2 function by direct phosphorylation and results in a dramatic decrease in the receptor-channel open probability (Po). We have identified S150 in the InsP3R2 core suppressor domain (amino acids 1–225) as the specific residue that is phosphorylated by CaMKII. Site-directed mutagenesis reveals that S150 is the CaMKII phosphorylation site responsible for modulation of channel activity. Nonphosphorylatable (S150A) and phosphomimetic (S150E) mutations were studied in planar lipid bilayers. The InsP3R2 S150A channel showed no decrease in activity when treated with CaMKII. Conversely, the phosphomimetic (S150E) channel displayed a very low Po under normal recording conditions in the absence of CaMKII (2 μm InsP3 and 250 nm [Ca2+]FREE) and mimicked a WT channel that has been phosphorylated by CaMKII. Phopho-specific antibodies demonstrate that InsP3R2 Ser-150 is phosphorylated in vivo by CaMKIIδ. The results of this study show that serine 150 of the InsP3R2 is phosphorylated by CaMKII and results in a decrease in the channel open probability.  相似文献   

7.
To investigate the effects of increasing concentrations ofmyo-inositol (inositol) on receptor stimulated [3H]inositol polyphosphate formation in the absence of lithium, slices of rat cerebral cortex were incubated with various concentrations of [3H]inositol (1 to 30 M). Carbachol stimulated formation of [3H]inositol trisphosphate (InsP3) and [3H]inositol 1,3,4,5-tetrakisphosphate {Ins(1,3,4,5)P4} increased several fold when the inositol concentration was increased reaching a plateau at approximately 12 M inositol. Time course studies revealed that in the presence of low concentrations of inositol (1 M), [3H]InsP3 and [3H]Ins(1,3,4,5)P4 formation in response to carbachol stimulation increased slowly over a 10 to 20 min time period, whereas in the presence of 4 and 12 M inositol, carbachol stimulated [3H]InsP3 and [3H]Ins(1,3,4,5)P4 formation was rapid and essentially complete within 3 to 5 min after carbachol addition. Although the carbachol dose response in 12 M inositol had a much greater maximal efficacy, there was no change in potency. Similar to the effects of carbachol on [3H]Ins(1,3,4,5)P4 formation from prelabeled phosphoinositides, muscarinic receptor stimulation increased Ins(1,3,4,5)P4 mass formation by seven fold. Furthermore, Li+ (8 mM) completely inhibited carbachol stimulated increases in Ins(1,3,4,5)P4 mass formation. In contrast to the effects of increasing inositol on carbachol stimulated formation of radiolabeled inositol phosphates, increasing inositol had no effect upon mass formation of Ins(1,3,4,5)P4. These results show that when measuring inositol polyphosphate formation by the radiolabeling technique in the absence of Li+, increasing the inositol concentration greatly increases the stimulated component of [3H]InsP3 and [3H]Ins(1,3,4,5)P4 formation. However, this inositol induced increase in agonist stimulated Ins(1,3,4,5)P4 formation is not reflected as an increase in mass formation.  相似文献   

8.
Abstract: Bovine adrenal chromaffin cells (BCC) were used to compare histamine- and angiotensin II-induced changes of inositol mono-, bis-, and trisphosphate (InsP1, InsP2, and InsP3, respectively) isomers, intracellular free Ca2+ ([Ca2+]i), and the pathways of inositol phosphate metabolism. Both agonists elevated [Ca2+]i by 200 nM 3–4 s after addition, but afterwards the histamine response was much more prolonged. Histamine and angiotensin II also produced similar four- to fivefold increases of Ins(1,4,5)P3 that peaked within 5 s. Over the first minute of stimulation, however, Ins(1,4,5)P3 formation was monophasic after angiotensin II, but biphasic after histamine, evidence supporting differential regulation of angiotensin II- and histamine-stimulated signal transduction. The metabolism of Ins(1,4,5)P3 by BCC homogenates was found to proceed via (a) sequential dephosphorylation to Ins(1,4)P2 and Ins(4)P, and (b) phosphorylation to inositol 1,3,4,5-tetrakisphosphate, followed by dephosphorylation to Ins(1,3,4)P3, Ins(1,3)P2, and Ins(3,4)P2, and finally to Ins(1 or 3)P. In whole cells, Ins(1 or 3)P only increased after histamine treatment. Additionally, Ins(1,3)P2 was the only other InsP2 besides Ins(1,4)P2 to accumulate within 1 min of agonist treatment [Ins(3,4)P2 did not increase]. These results support a correlation between the time course of Ins(1,4,5)P3 formation and the time course of [Ca2+]i transients and illustrate that Ca2+-mobilizing agonists can produce distinguishable patterns of inositol phosphate formation and [Ca2+], changes in BCC. Different patterns of second-messenger formation are likely to be important in signal recognition and may encode agonist-specific information.  相似文献   

9.
Flores S  Smart CC 《Planta》2000,211(6):823-832
 In response to abscisic acid (ABA), the duckweed Spirodela polyrrhiza (L.) activates a developmental pathway that culminates in the formation of dormant structures known as turions. Levels of the mRNA encoding d-myo-inositol-3-phosphate synthase (EC.5.5.1.4) which converts glucose-6-phosphate to inositol-3-phosphate, increase early in response to ABA. In order to understand the role of this enzyme in turion formation, we have investigated changes in inositol metabolism in ABA-treated plants. Here, we show that ABA-treatment leads to a 3-fold increase in free inositol, which peaks 2 d after treatment. This increase is followed by sequential increases in inositol phosphates and in accumulation of inositol hexakisphosphate (InsP6), in particular. In addition, we observed an early increase in a novel inositol bisphosphate which is not directly on the pathway to InsP6. In control plants, we observed synthesis and turnover of both inositol pentakisphosphate and InsP6. Two compounds more polar than InsP6 (diphosphoinositol polyphosphates) were present in both ABA-treated and control plants. Together, this suggests that the role of InsP6 in plants may be more complex than simply that of a storage compound during dormancy. Received: 10 January 2000 / Accepted: 25 February 2000  相似文献   

10.
Binding sites specific for inositol 1,4,5-trisphosphate (InsP3) have been demonstrated in sarcoplasmic reticulum vesicles isolated from heart muscle. Scatchard analysis of a binding isotherm indicated a high as well as a low affinity binding site [1]. In this study a comparison was made between InsP3 binding to crude microsomal membranes prepared from rat heart atria and ventricles respectively. Results obtained showed a four-fold higher incidence of binding to atrial membranes. Furthermore, the receptor populations of the atria and ventricles behaved differently during conditions causing fluctuations in tissue InsP3 levels, viz. ischaemia, reperfusion and 1-adrenergic stimulation. Reperfusion, as well as phenylephrine stimulation, caused an increase in InsP3 levels associated with down-regulation of the ventricular InsP3 receptor population while binding to atrial binding sites was elevated. In the ventricular population this down-regulation was the result of a reduction in Bmax alone with no changes in the Kd values of the high- or the low-affinity binding sites. The reason(s) for the differential response of the atrial and ventricular InsP3 receptor populations to changes in InsP3 levels, remains to be established.  相似文献   

11.
The phospholipase C (PLC; EC 3.1.4.3) activity in isolated plasma membranes of light-grown wheat (Triticum aestivum L. cv. Prelude) leaves was investigated. The activity against the polyphosphoinositides was strongly dependent on Ca2+ and was affected by the anionic detergent deoxycholate (DOC). In the presence of 20 M Ca2+ the PLC activity preferred phosphatidylinositol 4,5-bisphosphate (PIP2) over phosphatidylinositol 4-monophosphate (PIP) as a substrate. Instead, with 1 mM Ca2+ the enzyme clearly favoured PIP. In addition, the PIP2-PLC activity was increased by Mg2+ and in the presence of GTP, guanosine 5-(-thio)-triphosphate as well as ATP, CTP, guanosine 5-diphosphate and guanosine 5-(-thio)-diphosphate. Further analysis showed that a molybdate-sensitive phosphatase activity catalysing the dephosphorylation of inositol 1,4,5-trisphosphate (Ins(1,4,5)P3) is also associated with the plasma-membrane vesicles. Dephosphorylation of Ins(1,4,5)P3 was reduced in the presence of GTP or by inclusion of the unspecific phosphatase inhibitor molybdate. The results indicate the presence of a PIP2-PLC activity and the presence of a molybdate-sensitive phosphatase activity in wheat plasma-membrane vesicles.Abbreviations DOC deoxycholate - IDPase inosine 5-diphosphatase - InsPs inositol phosphates, the numbering at the end indicates the number of phosphate residues and when their positions on the inositol ring are known they are indicated in parentheses, i.e. - Ins(1,4,5)P3 inositol 1,4,5-trisphosphate - PIP phosphatidylinositol 4-monophosphate - PIP2 phosphatidylinositol 4,5-bisphosphate - PLC phospholipase C This work was financially supported by grant from the Deutsche Forschungsgemeinschaft (DFG). M. C. Arz gratefully acknowledges the support of a Graduiertenstipendium des Landes Nordrhein-Westfalen (Germany). We wish to thank S. Laden and G.E. Grambow for assistance.  相似文献   

12.
Inositol phosphates are a large and diverse family of signalling molecules. While genetic studies have discovered important functions for them, the biochemistry behind these roles is often not fully characterized. A key obstacle in inositol phosphate research in mammalian cells has been the lack of straightforward techniques for their purification and analysis. Here we describe the ability of titanium dioxide (TiO2) beads to bind inositol phosphates. This discovery allowed the development of a new purification protocol that, coupled with gel analysis, permitted easy identification and quantification of InsP6 (phytate), its pyrophosphate derivatives InsP7 and InsP8, and the nucleotides ATP and GTP from cell or tissue extracts. Using this approach, InsP6, InsP7 and InsP8 were visualized in Dictyostelium extracts and a variety of mammalian cell lines and tissues, and the effects of metabolic perturbation on these were explored. TiO2 bead purification also enabled us to quantify InsP6 in human plasma and urine, which led to two distinct but related observations. Firstly, there is an active InsP6 phosphatase in human plasma, and secondly, InsP6 is undetectable in either fluid. These observations seriously question reports that InsP6 is present in human biofluids and the advisability of using InsP6 as a dietary supplement.  相似文献   

13.
Cell-free extracts of Methanobacterium thermoautotrophicum were found to catalyze ATP synthesis from an endogeneous substrate. Synthesis was stimulated under hydrogen atmosphere and inhibited by KCL (K i =150 mM). Comparison of the properties of a number of cell constituents showed the endogeneous substrate to be 2,3-diphosphoglycerate. The compound is converted into 3-phosphoglycerate, and via 2-phosphoglycerate and phosphoenolpyruvate into pyruvate, at which the latter reaction is linked with ATP synthesis.Abbreviations HS-CoM Coenzyme M, 2-mercaptoethanesulfonate - CH3S-CoM methylcoenzyme m, 2-(methylthio)ethanesulfonate - HS-HTP 7-mercaptoheptanoyl-l-threonine phosphate - CoM-SS-HTP the heterodisulfide of HS-CoM and HS-HTP - BCFE bolled cell-free extract - TES N-tris(hydroxymethyl)methyl-2-aminoethanesulfonate - HEPES N-2-hydroxyethylpiperazine-N-ethanesulfonic acid - PEP phosphoenolpyruvate - 2,3-DPG 2,3-diphosphoglycerate - cDPG cyclic 2,3-diphosphoglycerate - 3-PG 3-phosphoglycerate - 2-PG 2-phosphoglycerate  相似文献   

14.
This essay attempts to summarize some of the best evidence for the role of inositol trisphosphate as a second messenger in signal transduction processes. The following aspects are addressed in the essay: (a) The synthesis of inositol trisphosphate and other inositol lipids, (b) Receptor-phosphatidylinositol bisphosphate phospholipase C coupling and the N-ras protooncogene, (c) Inositol trisphosphate and intracellular calcium, (d) Cell growth and oncogenes, (e) Receptors linked to the phosphatidylinositol cycle, (f) Phototransduction and (g) Interactions between inositol trisphosphate and other second messengers.Abbreviations Cyclic AMP Adenosine 3,5-cyclic monophosphate - Cyclic GMP Guanosine 3,5-cyclic monophosphate - DG sn, 1,2-Diacylglycerol - EGF Epidermal growth factor - GDP Guanosine diphosphate - GTP Guanosine triphosphate - IP Inositol 1-monophosphate - IP2 Inositol 1,4-diphosphate - IP3 Inositol 1,4,5-trisphosphate - PA Phosphatidic acid - PDGF Platelet-derived growth factor - PI Phosphatidylinositol - PIP Phosphatidylinositol 4-monophosphate - PIP2 Phosphatidylinositol 4,5-bisphosphate - PIP3 Phosphatidylinositol 3,4,5-trisphosphate - PLC Phospholipase C  相似文献   

15.
A soluble extract from pea (Pisum sativum L.) roots, when incubated with ATP and inositol 1,4,5-trisphosphate, produced an inositol tetrakisphosphate. The chromatographic properties of this inositol tetrakisphosphate, and of the products formed by its chemical degradation, identify it as inositol 1,4,5,6-tetrakisphosphate. No evidence was obtained for a 3-phosphorylation of inositol 1,4,5-trisphosphate. The importance of these observations with respect to inositol phosphates and calcium signalling in higher plants, is discussed.Abbreviations HPLC high-performance liquid chromatography - Ins(1,4,5)P3 inositol 1,4,5-trisphosphate - InsP4 inositol tetrakisphosphate J.A.C. gratefully acknowledges support from the Agricultural and Food Research Council, U.K., Plant Molecular Biology Initiative.  相似文献   

16.
Abstract: Stimulation of muscarinic receptors expressed in SH-SY5Y human neuroblastoma cells resulted in a complex profile of inositol 1,4,5-trisphosphate (InsP3) accumulation, with a dramatic increase (six- to eightfold) over the first 10 s (the “peak” phase) and subsequently, from ~60 s onward, maintained at a lower but sustained level (the “plateau” phase). Chelation of extracellular Ca2+ with EGTA or inhibition of Ca2+ channels with Ni2+ showed that the plateau phase was dependent upon Ca2+ entry. Furthermore, use of thapsigargin and EGTA to discharge and sequester Ca2+ from intracellular stores revealed that Ca2+ from this source was capable of supporting the peak phase of the InsP3 response. Carbachol-stimulated phosphoinositidase C activity in permeabilized SH-SY5Y cells was also shown to be highly dependent on free Ca2+ concentration (20–100 nM) and suggests that under normal conditions, InsP3 formation is enhanced by increases in cytosolic free Ca2+ concentration that accompany muscarinic receptor activation. Measurement of carbachol-stimulated total inositol phosphate accumulation in the presence of Li+ indicated that the initial rate of phosphoinositide hydrolysis (from 0 to 30 s) was about fivefold greater than that from 30 to 300 s. This rapid but partial desensitization of receptor-mediated phosphoinositide hydrolysis provides strong evidence for the mechanism underlying the changes in InsP3 accumulation over this time. Because very similar data were obtained in Chinese hamster ovary cells transfected with human m3 receptor cDNA, we suggest that although increases in cytosolic free Ca2+ concentration amplify InsP3 formation during stimulation of m3 muscarinic receptors, the primary factor that governs the profile of InsP3 accumulation is rapid, but partial, desensitization. Such desensitization does not appear to be mediated by changes in cytosolic Ca2+ or protein kinase C activity.  相似文献   

17.
Although it is well-accepted that the phosphatidylinositol signalling transduction pathway, producing inositol-1,4,5-P3 (InsP3) and inositol-1,3,4,5-P4 (InsP4) as second messengers, functions in heart muscle, virtually nothing is known about the roles of the higher inositol polyphosphates such as inositolhexakisphosphate (InsP6). This study demonstrates that InSP6 has the ability to bind intracellularly, with different binding characteristics, to different myocardial membranes. Binding to purified sarcoplasmic reticulum (SR) membranes, purified sarcolemmal (SL) membranes as well as to viable mitochondria were characterized. Binding to all these membranes display high as well as low affinity binding sites, with differing affinities. Kd values of binding to SR were 32 and 383 nM, to SL 61 and 1312 nM, while those of mitochondrial binding were 230 and 2200 nM respectively.InsP4 binding was also investigated and displayed the following characteristics: to SR, one low affinity binding site (Kd = 203 nM) and to SL, a high as well as a low affinity binding site with Kd values of 41 and 2075 nM respectively. Presence of InsP3, the second messenger for SR calcium release, at concentrations of 1 nM, elevated the binding of InsP4 to SR and SL by a mean of 30% and 20% respectively.Fractionation of SR and SL membranes on sucrose density gradients, after solubilization with CHAPS, indicated that InsP6 bound to two separate protein peaks in both these membranes, while InsP4 bound to only one. In SR membranes, InsP4 bound preferentially to a protein separating at high sucrose density while it bound to a protein separating at low sucrose density in SL membranes.  相似文献   

18.
Protein kinase A (PKA) phosphorylation of inositol 1,4,5-trisphosphate receptors (InsP3Rs) represents a mechanism for shaping intracellular Ca2+ signals following a concomitant elevation in cAMP. Activation of PKA results in enhanced Ca2+ release in cells that express predominantly InsP3R2. PKA is known to phosphorylate InsP3R2, but the molecular determinants of this effect are not known. We have expressed mouse InsP3R2 in DT40-3KO cells that are devoid of endogenous InsP3R and examined the effects of PKA phosphorylation on this isoform in unambiguous isolation. Activation of PKA increased Ca2+ signals and augmented the single channel open probability of InsP3R2. A PKA phosphorylation site unique to the InsP3R2 was identified at Ser937. The enhancing effects of PKA activation on this isoform required the phosphorylation of Ser937, since replacing this residue with alanine eliminated the positive effects of PKA activation. These results provide a mechanism responsible for the enhanced Ca2+ signaling following PKA activation in cells that express predominantly InsP3R2.Hormones, neurotransmitters, and growth factors stimulate the production of InsP33 and Ca2+ signals in virtually all cell types (1). The ubiquitous nature of this mode of signaling dictates that this pathway does not exist in isolation; indeed, a multitude of additional signaling pathways can be activated simultaneously. A prime example of this type of “cross-talk” between independently activated signaling systems results from the parallel activation of cAMP and Ca2+ signaling pathways (2, 3). Interactions between these two systems occur in numerous distinct cell types with various physiological consequences (36). Given the central role of InsP3R in Ca2+ signaling, a major route of modulating the spatial and temporal features of Ca2+ signals following cAMP production is potentially through PKA phosphorylation of the InsP3R isoform(s) expressed in a particular cell type.There are three InsP3R isoforms (InsP3R1, InsP3R2, and InsP3R3) expressed to varying degrees in mammalian cells (7, 8). InsP3R1 is the major isoform expressed in the nervous system, but it is less abundant compared with other subtypes in non-neuronal tissues (8). Ca2+ release via InsP3R2 and InsP3R3 predominate in these tissues. InsP3R2 is the major InsP3R isoform in many cell types, including hepatocytes (7, 8), astrocytes (9, 10), cardiac myocytes (11), and exocrine acinar cells (8, 12). Activation of PKA has been demonstrated to enhance InsP3-induced Ca2+ signaling in hepatocytes (13) and parotid acinar cells (4, 14). Although PKA phosphorylation of InsP3R2 is a likely causal mechanism underlying these effects, the functional effects of phosphorylation have not been determined in cells unambiguously expressing InsP3R2 in isolation. Furthermore, the molecular determinants of PKA phosphorylation of this isoform are not known.PKA-mediated phosphorylation is an efficient means of transiently and reversibly regulating the activity of the InsP3R. InsP3R1 was identified as a major substrate of PKA in the brain prior to its identification as the InsP3R (15, 16). However, until recently, the functional consequences of phosphorylation were unresolved. Initial conflicting results were reported indicating that phosphoregulation of InsP3R1 could result in either inhibition or stimulation of receptor activity (16, 17). Mutagenic strategies were employed by our laboratory to clarify this discrepancy. These studies unequivocally assigned phosphorylation-dependent enhanced Ca2+ release and InsP3R1 activity at the single channel level, through phosphorylation at canonical PKA consensus motifs at Ser1589 and Ser1755. The sites responsible were also shown to be specific to the particular InsP3R1 splice variant (18). These data were also corroborated by replacing the relevant serines with glutamates in a strategy designed to construct “phosphomimetic” InsP3R1 by mimicking the negative charge added by phosphorylation (19, 20). Of particular note, however, although all three isoforms are substrates for PKA, neither of the sites phosphorylated by PKA in InsP3R1 are conserved in the other two isoforms (21). Recently, three distinct PKA phosphorylation sites were identified in InsP3R3 that were in different regions of the protein when compared with InsP3R1 (22). To date, no PKA phosphorylation sites have been identified in InsP3R2.Interactions between Ca2+ and cAMP signaling pathways are evident in exocrine acinar cells of the parotid salivary gland. In these cells, both signals are important mediators of fluid and protein secretion (23). Multiple components of the [Ca2+]i signaling pathway in these cells are potential substrates for modulation by PKA. Previous work from this laboratory established that activation of PKA potentiates muscarinic acetylcholine receptor-induced [Ca2+]i signaling in mouse and human parotid acinar cells (4, 24, 25). A likely mechanism to explain this effect is that PKA phosphorylation increases the activity of InsP3R expressed in these cells. Consistent with this idea, activation of PKA enhanced InsP3-induced Ca2+ release in permeabilized mouse parotid acinar cells and also resulted in the phosphorylation of InsP3R2 (4).Invariably, prior work examining the functional effects of PKA phosphorylation on InsP3R2 has been performed using cell types expressing multiple InsP3R isoforms. For example, AR4-2J cells are the preferred cell type for examining InsP3R2 in relative isolation, because this isoform constitutes more than 85% of the total InsP3R population (8). InsP3R1, however, contributes up to ∼12% of the total InsP3R in AR4-2J cells. An initial report using InsP3-mediated 45Ca2+ flux suggested that PKA activation increased InsP3R activity in AR4-2J cells (21). A similar conclusion was made in a later study, which documented the effects of PKA activation on agonist stimulated Ca2+ signals in AR4-2J cells (26). Any effects of phosphorylation observed in these experiments could plausibly have resulted from phosphorylation of the residual InsP3R1.Although PKA enhances InsP3-induced calcium release in cells expressing predominantly InsP3R2, including hepatocytes, parotid acinar cells, and AR4-2J cells (4, 13, 21, 26, 27), InsP3R2 is not phosphorylated at stoichiometric levels by PKA (21). This observation has called into question the physiological significance of PKA phosphorylation of InsP3R2 (28). The apparent low levels of InsP3R2 phosphorylation are clearly at odds with the augmented Ca2+ release observed in cells expressing predominantly this isoform. The equivocal nature of these findings probably stems from the fact that, to date, all of the studies demonstrating positive effects of PKA activation on Ca2+ release were conducted in cells that also express InsP3R1. The purpose of the current experiments was to analyze the functional effects of phosphorylation on InsP3R2 expressed in isolation on a null background. We report that InsP3R2 activity is increased by PKA phosphorylation under these conditions, and furthermore, we have identified a unique phosphorylation site in InsP3R2 at Ser937. In total, these results provide a direct mechanism for the cAMP-induced activation of InsP3R2 via PKA phosphorylation of InsP3R2.  相似文献   

19.
Emerging studies have suggested that there is a close link between inositol phosphate (InsP) metabolism and cellular phosphate (Pi) homeostasis in eukaryotes; however, whether a common InsP species is deployed as an evolutionarily conserved metabolic messenger to mediate Pi signaling remains unknown. Here, using genetics and InsP profiling combined with Pi‐starvation response (PSR) analysis in Arabidopsis thaliana, we showed that the kinase activity of inositol pentakisphosphate 2‐kinase (IPK1), an enzyme required for phytate (inositol hexakisphosphate; InsP6) synthesis, is indispensable for maintaining Pi homeostasis under Pi‐replete conditions, and inositol 1,3,4‐trisphosphate 5/6‐kinase 1 (ITPK1) plays an equivalent role. Although both ipk1‐1 and itpk1 mutants exhibited decreased levels of InsP6 and diphosphoinositol pentakisphosphate (PP‐InsP5; InsP7), disruption of another ITPK family enzyme, ITPK4, which correspondingly caused depletion of InsP6 and InsP7, did not display similar Pi‐related phenotypes, which precludes these InsP species from being effectors. Notably, the level of d /l ‐Ins(3,4,5,6)P4 was concurrently elevated in both ipk1‐1 and itpk1 mutants, which showed a specific correlation with the misregulated Pi phenotypes. However, the level of d /l ‐Ins(3,4,5,6)P4 is not responsive to Pi starvation that instead manifests a shoot‐specific increase in the InsP7 level. This study demonstrates a more nuanced picture of the intersection of InsP metabolism and Pi homeostasis and PSRs than has previously been elaborated, and additionally establishes intermediate steps to phytate biosynthesis in plant vegetative tissues.  相似文献   

20.
In order to test the hypothesis that excitation in Drosophila photoreceptors is mediated by Ca2+ released from internal stores, the Ca2+ buffers EGTA, BAPTA and di-bromo-BAPTA (DBB) were introduced into dissociated photoreceptors via whole-cell recording pipettes. All buffers were preloaded with Ca2+ to provide the same free Ca2+ concentration (250 nM). EGTA (up to 18 mM free buffer) had only weak effects upon voltage-clamped flash responses in normal Ringer's solution (1.5 mM Ca 0 2+ ), and no effect in Ca2+-free solution. The maximum BAPTA concentration tested (14.4 mM free BAPTA) reduced the initial rate of rise by ca. 5000-fold in normal Ringer's solution; by ca. 500-fold in Ca2+free solution; and only ca. 60-fold in the absence of Mg2+, which preferentially blocks one component of the light-sensitive current. Although BAPTA delayed the time-to-peak in normal Ringer's solution, responses in Ca2+ free Ringer's solution were accelerated. These results support the role of Ca2+ influx in regulating sensitivity and response kinetics; however, in view of the high concentrations required to attenuate responses in Ca2+ free Ringer's solution, the role of Ca2+ release in excitation remains unclear. DBB was ca. 2–3 fold more potent than BAPTA, and at concentrations > 5 mM had a qualitatively different action, greatly delaying the time-to-peak. This suggests DBB may have distinct pharmacological actions or access to compartments inaccessible to BAPTA.The only current activated by introducing 5–500 M Ca2+ (buffered with nitrilo-triacetic acid) was electrogenic Na+/Ca2+ exchange. When this was blocked by removing Nao 0 + , a novel cationic conductance was activated. However, its properties did not resemble those the light-activated conductance, and thus do not support the hypothesis that Ca2+ is sufficient for excitation.Abbreviations BAPTA bis-(o-aminophenoxy)-ethane-N,N,N-tetracetic acid - DBB Di-bromo-bapta - NTA nitrilo-triacetic acid - InsP 3 inositol 1,4,5-trisphosphate  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