首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
The yeast and Drosophila pathways leading to the production of inositol hexakisphosphate (InsP(6)) have been elucidated recently. The in vivo pathway in humans has been assumed to be similar. Here we show that overexpression of Ins(1,3,4)P(3) 5/6-kinase in human cell lines results in an increase of inositol tetrakisphosphate (InsP(4)) isomers, inositol pentakisphosphate (InsP(5)) and InsP(6), whereas its depletion by RNA interference decreases the amounts of these inositol phosphates. Expression of Ins(1,3,4,6)P(4) 5-kinase does not increase the amount of InsP(5) and InsP(6), although its depletion does block InsP(5) and InsP(6) production, showing that it is necessary for production of InsP(5) and InsP(6). Expression of Ins(1,3,4,5,6)P(5) 2-kinase increases the amount of InsP(6) by depleting the InsP(5) in the cell, and depletion of 2-kinase decreases the amount of InsP(6) and causes an increase in InsP(5). These results are consistent with a pathway that produces InsP(6) through the sequential action of Ins(1,3,4)P(3) 5/6-kinase, Ins(1,3,4,6)P(4) 5-kinase, and Ins(1,3,4,5,6)P5 2-kinase to convert Ins(1,3,4)P(3) to InsP(6). Furthermore, the evidence implicates 5/6-kinase as the rate-limiting enzyme in this pathway.  相似文献   

3.
It is important to both physiological and pathological osteogenesis to understand the significance of changes in gene expression in growth-plate chondrocytes that transit between the proliferative and hypertrophic states. MINPP is one such gene of interest. The Minpp protein dephosphorylates highly phosphorylated inositol signaling molecules InsP(5) and InsP(6). We show here that the ATDC5 chondrocyte progenitor cell line can recapitulate developmentally specific changes in MINPP expression previously only seen in longitudinal bone growth plates-both an initial 2-3-fold increase and a subsequent decrease back to initial levels during transition to hypertrophy. The increase in MINPP expression was accompanied by a 40% decrease in InsP(6) levels in ATDC5 cells. However, InsP(5) levels were not modified. Furthermore, throughout the hypertrophic phase, during which MINPP expression decreased, there were no alterations in InsP(5) and InsP(6) levels. We also created an ATDC5 line that stably overexpressed Minpp at 2-fold higher levels than in wild-type cells. This had no significant effect upon cellular levels of InsP(5) and InsP(6). Thus, substantial changes in MINPP expression can occur without a net effect upon InsP(5) and InsP(6) turnover in vivo. On the other hand, Minpp-overexpressing cells showed impaired chondrogenesis. We noted that the expression of alkaline phosphatase activity was inversely correlated with the expression of MINPP. The ATDC5 cells that overexpress Minpp failed to show an insulin-dependent increase in alkaline phosphatase levels, which presumably affects phosphate balance [J. Biol. Chem. 276 (2001) 33995], and may be the reason cellular differentiation was impaired. In any case, we conclude that Minpp is important to chondrocyte differentiation, but in a manner that is, surprisingly, independent of inositol polyphosphate turnover.  相似文献   

4.
Rat 6 fibroblasts that stably overexpress cDNA for the beta 1 isozyme of protein kinase C (PKC3 cells) were used to determine the effect of protein kinase C (PKC) overexpression on hormonal stimulation of phospholipid hydrolysis. In control Rat 6 cells, inositol trisphosphate levels (InsP3) were increased 9-fold in 15 s in response to 10 nM alpha-thrombin, compared with only a 2-fold increase in PKC3 cells. PKC overexpression also inhibited thrombin-stimulated production of 1,2-diacylglycerol, the other product of phosphatidylinositol 4,5-bisphosphate hydrolysis, by 73% at 15 s. In permeabilized cells, PKC overexpression greatly reduced guanosine thiotriphosphate-stimulated InsP3 accumulation, but did not affect InsP3 stimulation by increased free calcium concentration. These data suggest that desensitization of thrombin-stimulated phosphoinositide-phospholipase C is enhanced by PKC-beta 1 overexpression and may involve modulation of G-protein/phospholipase C coupling. In contrast, thrombin was 4.5-fold more effective in stimulation of phosphatidylcholine-phospholipase D activity in PKC3 cells than in control cells, as determined by phosphatidylethanol formation. In permeabilized cells, guanosine thiotriphosphate also stimulated phospholipase D activity more effectively in PKC3 cells than in control cells, suggesting that upregulation of phospholipase D activity by PKC overexpression occurs distal to the thrombin receptor. These results suggest that PKC may act as a switch to up-regulate phosphatidylcholine-phospholipase D and down-regulate phosphoinositide-phospholipase C stimulations.  相似文献   

5.
NIH 3T3 fibroblasts were stably transfected with rat brain inositol 1,4,5-trisphosphate (Ins(1,4,5)P3) 3-kinase to explore the relationship between increased production of Ins(1,3,4,5)P4 and the formation of InsP5 and InsP6. Mass measurements of InsP5 and InsP6 revealed no significant difference between kinase- and vector-transfected fibroblasts. However, such 3-kinase-transfected cells, when labeled with [3H]inositol for 48-72 h, showed lower levels of [3H]InsP5 and [3H]InsP6, as well as [3H]Ins(1,3,4,6)P4 and D/L[3H]Ins(1,4,5,6)P4, than their vector-transfected counterparts. Because Ins(1,4,5)P3 3-kinase-transfected cells grew less rapidly than vector-transfected controls, we determined whether the synthesis of InsP5 and InsP6 was related to a specific phase of the cell cycle. When NIH 3T3 cells prelabeled with [3H]inositol were synchronized by serum deprivation followed by stimulation with platelet-derived growth factor (PDGF), the amounts of labeled InsP5 and InsP6 began to increase only after 12 h of stimulation, when cells entered the S-phase as indicated by increased [3H]thymidine incorporation. The enhanced synthesis of these inositol polyphosphates was preceded by an early increase in Ins(1,4,5)P3 and its metabolites that was no longer evident by the fifth hour of PDGF action. There was also a prominent and biphasic increase in the level of D/L-Ins(1,4,5,6)P4 with an early peak at approximately 3 h and a second rise that paralleled the increases in InsP5 and InsP6. These results indicate that the formation of highly phosphorylated inositols is not tightly coupled to the receptor-mediated formation of Ins(1,4,5)P3 and its metabolites but is mainly determined by other factors that operate at specific points of the cell cycle.  相似文献   

6.
These studies explore the molecular effect of arsenicals on MM cells. Freshly isolated cells derived from patients with advanced, chemo-refractory myeloma as well as human myeloma cell lines, ARP-1, RPMI-8226 and H929 were exposed to the organic arsenical melarsoprol and to the inorganic compound AT. Both agents potently induced apoptosis in myeloma cells. Exposure to 1-5 microM AT or melarsoprol for 6 hours suppressed NF-kappa B DNA binding and enhanced of c-Jun kinase (JNK) activity. Arsenic also activated caspase-3 resulting in the cleavage of poly (ADP-ribose) polymerase (PARP) and Fas/TNF alpha related receptor interacting protein (RIP). In contrast to reported observations in acute promyelocytic leukemia, myeloma cell apoptosis was not associated with either the downregulation of Bcl-2 protein or with alterations in the expression of other Bcl-2 family members, Bax, Bak, Bag, and Bcl-xl. This study first shows that arsenic induces apoptotic signaling in MM through the cleavage of TNF alpha related receptor interacting protein (RIP). RIP is a key downstream protein in FasL/ TNF alpha /TRAIL induced apoptosis and a major antiapoptotic adaptor of pathways through NF-kappa B and JNK. RIP has not been previously characterized in myeloma. This study supports the hypothesis that arsenicals share common mediators (RIP, NF-kappa B, PARP, caspase-3) with death receptor induced apoptosis. These studies provide an important insight into the molecular mechanism of AT induced apoptosis and can be used in the development of adjuvant therapy for MM, presently an incurable disease.  相似文献   

7.
The enzyme(s) responsible for the production of inositol hexakisphosphate (InsP(6)) in vertebrate cells are unknown. In fungal cells, a 2-kinase designated Ipk1 is responsible for synthesis of InsP(6) by phosphorylation of inositol 1,3,4,5,6-pentakisphosphate (InsP(5)). Based on limited conserved sequence motifs among five Ipk1 proteins from different fungal species, we have identified a human genomic DNA sequence on chromosome 9 that encodes human inositol 1,3,4,5,6-pentakisphosphate 2-kinase (InsP(5) 2-kinase). Recombinant human enzyme was produced in Sf21 cells, purified, and shown to catalyze the synthesis of InsP(6) or phytic acid in vitro. The recombinant protein converted 31 nmol of InsP(5) to InsP(6)/min/mg of protein (V(max)). The Michaelis-Menten constant for InsP(5) was 0.4 microM and for ATP was 21 microM. Saccharomyces cerevisiae lacking IPK1 do not produce InsP(6) and show lethality in combination with a gle1 mutant allele. Here we show that expression of the human InsP(5) 2-kinase in a yeast ipk1 null strain restored the synthesis of InsP(6) and rescued the gle1-2 ipk1-4 lethal phenotype. Northern analysis on human tissues showed expression of the human InsP(5) 2-kinase mRNA predominantly in brain, heart, placenta, and testis. The isolation of the gene responsible for InsP(6) synthesis in mammalian cells will allow for further studies of the InsP(6) signaling functions.  相似文献   

8.
Phospholipase C cleaves phosphatidylinositol 4,5-bisphosphate to form both inositol 1,4,5-trisphosphate (Ins(1,4,5)P3) and inositol 1,2-cyclic 4,5-trisphosphate (cInsP3). The further metabolism of these inositol trisphosphates is determined by two enzymes: a 3-kinase and a 5-phosphomonoesterase. The first enzyme converts Ins(1,4,5)P3 to inositol 1,3,4,5-tetrakisphosphate (InsP4), while the latter forms inositol 1,4-bisphosphate and inositol 1,2-cyclic 4-bisphosphate from Ins(1,4,5)P3 and cInsP3, respectively. The current studies show that the 3-kinase is unable to phosphorylate cInsP3. Also, the 5-phosphomonoesterase hydrolyzes InsP4 with an apparent Km of 0.5-1.0 microM to form inositol 1,3,4-trisphosphate at a maximal velocity approximately 1/30 that for Ins(1,4,5)P3. The apparent affinity of the enzyme for the three substrates is InsP4 greater than Ins(1,4,5)P3 greater than cInsP3; however, the rate at which the phosphatase hydrolyzes these substrates is Ins(1,4,5)P3 greater than cInsP3 greater than InsP4. The 5-phosphomonoesterase and 3-kinase enzymes may control the levels of inositol trisphosphates in stimulated cells. The 3-kinase has a low apparent Km for Ins(1,4,5)P3 as does the 5-phosphomonoesterase for InsP4, implying that the formation and breakdown of InsP4 may proceed when both it and its precursor are present at low levels. Ins(1,4,5)P3 is utilized by both the 3-kinase and 5-phosphomonoesterase, while cInsP3 is utilized relatively poorly only by the 5-phosphomonoesterase. These findings imply that inositol cyclic trisphosphate may be metabolized slowly after its formation in stimulated cells.  相似文献   

9.
10.
The inhibitor of apoptosis (IAP) proteins are a family of anti-apoptotic regulators found in viruses and metazoans. c-IAP1 and c-IAP2 are recruited to tumor necrosis factor receptor 1 (TNFR1)-associated complexes where they can regulate receptor-mediated signaling. Both c-IAP1 and c-IAP2 have been implicated in TNFalpha-stimulated NF-kappaB activation. However, individual c-IAP1 and c-IAP2 gene knock-outs in mice did not reveal changes in TNF signaling pathways, and the phenotype of a combined deficiency of c-IAPs has yet to be reported. Here we investigate the role of c-IAP1 and c-IAP2 in TNFalpha-stimulated activation of NF-kappaB. We demonstrate that TNFalpha-induced NF-kappaB activation is severely diminished in the absence of both c-IAP proteins. In addition, combined absence of c-IAP1 and c-IAP2 rendered cells sensitive to TNFalpha-induced cell death. Using cells with genetic ablation of c-IAP1 or cells where the c-IAP proteins were eliminated using IAP antagonists, we show that TNFalpha-induced RIP1 ubiquitination is abrogated in the absence of c-IAPs. Furthermore, we reconstitute the ubiquitination process with purified components in vitro and demonstrate that c-IAP1, in collaboration with the ubiquitin conjugating enzyme (E2) enzyme UbcH5a, mediates polymerization of Lys-63-linked chains on RIP1. Therefore, c-IAP1 and c-IAP2 are required for TNFalpha-stimulated RIP1 ubiquitination and NF-kappaB activation.  相似文献   

11.
The distribution of inositol 1,4,5-trisphosphate (InsP3) 3-kinase mRNA in the rat brain is reported using oligonucleotides based on a cDNA clone sequence that encodes rat brain InsP3 3-kinase and the in situ hybridization technique. Moderate levels were found in CA2-4 pyramidal neurons, in the cortex, and in the striatum. The cerebellar granule cells, thalamus, hypothalamus, brainstem, spinal cord, and white matter tracts were almost negative. The levels of InsP3 3-kinase mRNA were highest in the hippocampal CA1 pyramidal neurons, granule cells of the dentate gyrus, and cerebellar Purkinje cells. These results contrast with the lower concentration of the InsP3 receptor already reported in the hippocampus versus the Purkinje cells and suggest a special role for inositol 1,3,4,5-tetrakisphosphate in Ammon's horn.  相似文献   

12.
It has been proposed that the inositol 1,4,5-trisphosphate receptor (InsP(3)R) type III acts as a trigger for InsP(3)-mediated calcium (Ca(2+)) signaling, because this InsP(3) isoform lacks feedback inhibition by cytosolic Ca(2+). We tested this hypothesis in RIN-m5F cells, which express predominantly the type III receptor. Extracellular ATP increases Ca(2+) in these cells, and we found that this effect is independent of extracellular Ca(2+) but is blocked by the InsP(3)R antagonist heparin. There was a dose-dependent increase in the number of cells responding to ATP and two-photon flash photolysis of caged-Ca(2+) heightened the sensitivity of RIN-m5F cells to this increase. These findings provide evidence that Ca(2+) increases the sensitivity of the InsP(3)R type III in intact cells and supports the idea that this isoform can act as a trigger for hormone-induced Ca(2+) signaling.  相似文献   

13.
Stimulation of NIH-3T3 cells with prostaglandin F2 alpha (PGF2 alpha) caused a dose- and time-dependent generation of inositol phosphates. The first detectable changes were in the levels of Ins(1,4,5)P3 and Ins(1,3,4,5)P4. Increases in Ins(1,3,4)P3, InsP2 and InsP were detected later, and only minor changes were observed in putative InsP5 or InsP6. The accumulation of inositol phosphates was synergistically increased by the addition of calf serum, whereas PGF2 alpha had no effects on cell proliferation in either the presence or the absence of calf serum. Stimulation of a different clone of NIH-3T3 cells (AmNIH-3T3) or Swiss 3T3 cells with PGF2 alpha resulted in both inositol phospholipid breakdown and cell proliferation. No differences were found in the characteristics of PGF2 alpha-stimulated inositol phosphate generation between the two clones of NIH-3T3 cells, nor was there any difference in receptor number of Kd. These results question the role of inositol phospholipid breakdown in mitogenesis and demonstrate significant differences in the biochemical properties of apparently the 'same' cells.  相似文献   

14.
The transforming protein of polyomavirus, middle T (mT), forms a complex with two cellular enzymes: the protein tyrosine kinase pp60c-src and a phosphatidylinositol (PtdIns) 3-kinase. A mutant virus, Py1178T, encodes an mT protein which associates with and activates pp60c-src to the same extent as the wild type but fails to associate with PtdIns 3-kinase. To investigate relationships between activation of pp60c-src, association of PtdIns 3-kinase, and cellular levels of the second messenger inositol 1,4,5-trisphosphate (InsP3), we examined the effects of wild-type and mutant mT proteins on inositol metabolism in rat and mouse fibroblasts. Expression of either wild-type or 1178T mT caused a 300 to 500% increase in the InsP3 level. Cells transformed by Rous sarcoma virus also showed similar increases in InsP3 levels. Mutant mT proteins which failed to activate pp60c-src (NG59 and 1387T) had no effect on InsP3 levels. Pulse-chase experiments with [3H]inositol showed that the turnover of phosphoinositides was increased in cells transformed by either wild-type polyomavirus or Py1178T as compared with the normal parent cell line. The turnover of inositol phosphates was unchanged upon transformation. These data indicate that cells expressing either wild-type or mutant 1178T mT or pp60v-src exhibit elevated levels of InsP3 because of activation of phospholipase C. This activation appears to depend, directly or indirectly, upon activation of pp60src protein kinase activity. Activation of pp60c-src and elevation of InsP3 content are not sufficient for full transformation. Full transformation also requires the association of mT-pp60c-src complexes with PtdIns 3-kinase.  相似文献   

15.
A cDNA cloned from its ability to stimulate inorganic phosphate uptake in Xenopus oocytes (phosphate uptake stimulator (PiUS)) shows significant similarity with inositol 1,4,5-trisphosphate 3-kinase. However, the expressed PiUS protein showed no detectable activity against inositol 1,4,5-trisphosphate, nor the 1,3,4,5- or 3,4,5, 6-isomers of inositol tetrakisphosphate, whereas it was very active in converting inositol hexakisphosphate (InsP(6)) to inositol heptakisphosphate (InsP(7)). PiUS is a member of a family of enzymes found in many eukaryotes and we discuss the implications of this for the functions of InsP(7) and for the evolution of inositol phosphate kinases.  相似文献   

16.
Inositol hexakisphosphate (InsP6) increased 45Ca2+ uptake in cultured cerebellar granule cells. This increase was concentration dependent (EC50 = 20 microM), exhibited slow kinetics, and was present after 5 days of cell maturation in culture. InsP6 also enhanced D-[3H]aspartate release in cerebellar granule cells at 11-12 days in vitro. Stimulation of 45Ca2+ uptake was also produced by inositol pentakisphosphate but not by inositol 1,3,4,5-tetrakisphosphate. The increase in 45Ca2+ influx induced by InsP6 was independent of extracellular Na+ and was only partially reduced by the organic calcium channel blocker nifedipine. The intrinsic action of InsP6 was not affected by competitive or noncompetitive glutamate receptor antagonists. In addition, stimulations of 45Ca2+ uptake by InsP6 and glutamate were additive. These data provide evidence that InsP6 directly activates a specific population of neurons in the CNS.  相似文献   

17.
Wang Y  Sun X  Wu J  Xu BE  Gu C  Wang H  Wang X  Tan F  Peng X  Qiang B  Yuan J  Luo Y 《Biochemistry》2008,47(1):441-448
Tumor necrosis factor alpha (TNFalpha) triggers a signaling pathway converging on the activation of NF-kappaB, which forms the basis for many physiological and pathological processes. In a kinase gene screen using a NF-kappaB reporter, we observed that overexpression of casein kinase 1alpha (CK1alpha) enhanced TNFalpha-induced NF-kappaB activation, and a CK1alpha kinase dead mutant, CK1alpha (K46A), reduced NF-kappaB activation induced by TNFalpha. We subsequently demonstrated that CK1alpha interacted with receptor interacting protein 1 (RIP1) but not with TRADD, TRAF2, MEKK3, IKKalpha, IKKbeta, or IKKgamma in mammalian cells. RIP1 is an indispensable molecule in TNFalpha/NF-kappaB signaling. We demonstrated that CK1alpha interacted with and phosphorylated RIP1 at the intermediate domain. Finally, we showed that CK1alpha enhanced RIP1-mediated NF-kappaB activation. Taken together, our studies suggest that CK1alpha is another kinase that regulates RIP1 function in NF-kappaB activation.  相似文献   

18.
Associated with the middle T antigen of polyomavirus is a novel phosphatidylinositol (PtdIns) kinase activity which phosphorylates PtdIns at the D-3 position of the inositol ring. We have undertaken an analysis of myo-[3H]inositol-containing compounds in a panel of NIH 3T3 cell lines stably transfected with transforming and nontransforming middle T antigen mutants. All cell lines from which PtdIns 3-kinase activity coprecipitated with middle T antigen exhibited modestly elevated levels of PtdIns(3)P and compounds with predicted PtdIns(3,4)P2 and PtdIns(3,4,5)P3 structures. Complex formation between middle T antigen and PtdIns 3-kinase correlated not with an increase in total inositol phosphate levels but rather with elevated levels of InsP2 and InsP4. A specific increase in the level of an InsP2 species which comigrated in high-pressure liquid chromatography analysis with Ins(3,4)P2 was observed. These results suggest that association of the polyomavirus middle T antigen with PtdIns 3-kinase activates a distinct inositol metabolic pathway.  相似文献   

19.
We have demonstrated that the human homolog of the rat inositol phosphate multikinase is an inositol 1,3,4,6-tetrakisphosphate 5-kinase (InsP(4) 5-kinase). The cDNA of the human gene contained a putative open reading frame of 1251 bp encoding 416 amino acids with 83.6% identity compared with the rat protein. The substrate specificity of the recombinant human protein demonstrated preference for Ins(1,3,4,6)P(4) with a catalytic efficiency (V(max)/K(m)) 43-fold greater than that of Ins(1,3,4,5)P(4) and 2-fold greater than that of Ins(1,4,5)P(3). The apparent V(max) was 114 nmol of Ins(1,3,4,5,6)P(5) formed/min/mg of protein, and the apparent K(m) was 0.3 microm Ins(1,3,4,6)P(4). The functional homolog in yeast is Ipk2p, and ipk2-null yeast strains do not synthesize Ins(1,3,4,5,6)P(5) or InsP(6). Synthesis of these compounds was restored by transformation with wild-type yeast IPK2 but not with human InsP(4) 5-kinase. Thus the human gene does not complement for the loss of the yeast gene because yeast cells do not contain the substrate Ins(1,3,4,6)P(4), and the reaction of the human protein with Ins(1,3,4,5)P(4) is insufficient to effect rescue or synthesis of InsP(5) and InsP(6). Therefore the major activity of human InsP(4) 5-kinase is phosphorylation at the D-5 position, and the pathways for synthesis of Ins(1,3,4,5,6)P(5) in yeast versus humans are different.  相似文献   

20.
To try to further define the mechanism of action of the putative second messenger inositol 1,3,4,5-tetrakisphosphate (InsP4), we have studied its effects in permeabilized cells expressing different levels of inositol trisphosphate receptor (InsP3R) types I and III and of the GTPase-activating protein GAP1IP4BP. During the growth curve of human HL-60 cells and mouse T15 cells there was an increase in these proteins, which was further increased by differentiation (HL-60) and, marginally, by transformation (T15). T15 cells entering the stationary phase showed much lower concentrations of these proteins and expression was below detection in apoptotic HL-60 cells. Rasp21 showed a different pattern of expression. The ratios of InsP3R subtypes seem to affect the dose-response curve for inositol 2,4,5-trisphosphate Ins(2,4,5)P3. In permeabilized T15 cells the curve was approximately 5-fold to the right of that obtained using HL-60 cells. However, permeabilized untreated and differentiated HL-60 cells and T15 cells all showed a comparable synergistic effect of InsP4 on Ca2+ release stimulated by a concentration of Ins(2,4,5)P3, releasing approximately 20% of the Ins(1,4,5)P3 sensitive Ca2+ pool. The data indicate that under these conditions InsP4 is acting independently of cell type, of the ratio of inositol trisphosphate receptor subtypes, and of the concentration of GAP1IP4BP.  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