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 203 毫秒
1.
To understand how a signaling molecule's activities are regulated, we need insight into the processes controlling the dynamic balance between its synthesis and degradation. For the Ins(1,3,4,5,6)P5 signal, this information is woefully inadequate. For example, the only known cytosolic enzyme with the capacity to degrade Ins(1,3,4,5,6)P5 is the tumour-suppressor PTEN [J.J. Caffrey, T. Darden, M.R. Wenk, S.B. Shears, FEBS Lett. 499 (2001) 6 ], but the biological relevance has been questioned by others [E.A. Orchiston, D. Bennett, N.R. Leslie, R.G. Clarke, L. Winward, C.P. Downes, S.T. Safrany, J. Biol. Chem. 279 (2004) 1116 ]. The current study emphasizes the role of physiological levels of PTEN in Ins(1,3,4,5,6)P5 homeostasis. We employed two cell models. First, we used a human U87MG glioblastoma PTEN-null cell line that hosts an ecdysone-inducible PTEN expression system. Second, the human H1299 bronchial cell line, in which PTEN is hypomorphic due to promoter methylation, has been stably transfected with physiologically relevant levels of PTEN. In both models, a novel consequence of PTEN expression was to increase Ins(1,3,4,5,6)P5 pool size by 30-40% (p<0.01); this response was wortmannin-insensitive and, therefore, independent of the PtdIns 3-kinase pathway. In U87MG cells, induction of the G129R catalytically inactive PTEN mutant did not affect Ins(1,3,4,5,6)P(5) levels. PTEN induction did not alter the expression of enzymes participating in Ins(1,3,4,5,6)P5 synthesis. Another effect of PTEN expression in U87MG cells was to decrease InsP6 levels by 13% (p<0.02). The InsP6-phosphatase, MIPP, may be responsible for the latter effect; we show that recombinant human MIPP dephosphorylates InsP6 to D/L-Ins(1,2,4,5,6)P5, levels of which increased 60% (p<0.05) following PTEN expression in U87MG cells. Overall, our data add higher inositol phosphates to the list of important cellular regulators [Y. Huang, R.P. Wernyj, D.D. Norton, P. Precht, M.C. Seminario, R.L. Wange, Oncogene, 24 (2005) 3819 ] the levels of which are modulated by expression of the highly pleiotropic PTEN protein.  相似文献   

2.
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   

3.
Three myo-inositol tetrakisphosphate analogues were synthesised based upon myo-inositol 1,3,4,6-tetrakisphosphate: 2,5-di-O-methyl myo-inositol-1,3,4,6-tetrakisphosphate 19 and its phosphorothioate derivative 22, together with myo-inositol 1,3,4,6 tetrakisphosphorothioate 25. These compounds were prepared by phosphitylating 2,5-di-O-methyl-myo-inositol and 2,5-di-O-benzyl-myo-inositol followed by oxidation with t-butylhydroperoxide or sulfoxidation at room temperature using sulfur in a mixed solvent of DMF and pyridine. Sulfoxidation was complete within 15 min; however, without DMF, the reaction was much slower, and required overnight. When evaluated against Ins(1,4,5)P(3) 5-phosphatase, 3-kinase and for Ca(2+) release at the Ins(1,4,5)P(3) receptor, only weak activity was observed for Ca(2+) release. 22 and 25 are potent 5-phosphatase inhibitors and 25 is a moderate inhibitor of 3-kinase. Thus, we have synthesised potent enzyme inhibitors, which do not mobilise Ca(2+) and devised conditions for quick, clean and inexpensive sulfoxidation of inositol polyphosphite intermediates.  相似文献   

4.
Inositol 1,3,4-trisphosphate (Ins(1,3,4)P3) metabolism has been studied in liver homogenates and in 100,000 x g supernatant and particulate fractions. When liver homogenates were incubated in an "intracellular" medium containing 5 mM MgATP, equal proportions of Ins(1,3,4)P3 were dephosphorylated and phosphorylated. Two inositol tetrakisphosphate (InsP4) products and an inositol pentakisphosphate (InsP5) were detected. The InsP4 isomers were unequivocally identified as inositol 1,3,4,5-tetrakisphosphate (Ins(1,3,4,5)P4) and inositol 1,3,4,6-tetrakisphosphate (Ins(1,3,4,6)P4) by high performance liquid chromatography separation of inositol phosphates, periodate oxidation, alkaline hydrolysis, and stereo-specific polyol dehydrogenase. Ins(1,3,4)P3 5-kinase is a novel enzyme activity and accounted for 16% of the total Ins(1,3,4)P3 phosphorylation. Ins(1,3,4,6)P4 was also shown to be further phosphorylated to inositol 1,3,4,5,6-pentakisphosphate (Ins(1,3,4,5,6)P5) by a kinase not previously known to occur in liver. About 75% of Ins(1,3,4)P3 kinase activities were soluble and were partly purified by anion-exchange fast protein liquid chromatography. The two Ins(1,3,4)P3 kinase activities eluted as a single peak that was well resolved from Ins(1,3,4)P3 phosphatase, Ins(1,3,4,6)P4 5-kinase, and Ins(1,3,4,5)P4 5-phosphatase activities. A further novel observation was that 10 microM Ins(1,3,4,5)P4 inhibited Ins(1,3,4)P3 kinase activities by 60%.  相似文献   

5.
The metabolism of biologically active inositol phosphates in developed ovarian follicles from Xenopus laevis was investigated. Techniques used were microinjection of tracer into the intact oocyte coupled by gap junctions to follicle cells, as well as addition of tracer to homogenates of ovarian follicles and to homogenates of oocytes stripped of outer follicle-cell layers. Metabolism was similar to that previously described for other types of cell and tissue, with several unusual features. Homogenates of ovarian follicles were shown to contain an apparent 3'-phosphomonoesterase capable of converting [3H]Ins(1,3,4,5)P4 predominantly into a substance with h.p.l.c. elution characteristics of Ins(1,4,5)P3. In intact ovarian follicles, little Ins(1,4,5)P3 was formed but the esterase was activated by the phorbol ester activator of protein kinase C, PMA (phorbol 12-myristate 13-acetate; 60 nM), as well as by acetylcholine (200 microM). In follicle homogenates, this enzyme also appeared to be active in converting [3H]Ins(1,3,4)P3 into a substance eluting as Ins(1,4)P2. The apparent 3'-phosphomonoesterase activity was not inhibited by intracellular (or higher) levels of Mg2+. Although PMA activated this enzyme in intact oocytes relative to 5'-phosphomonoesterase activation, it did not enhance overall metabolism, in contrast with reports on other tissues. Compared with the processing of inositol phosphates injected into the intact follicle, homogenization in simulated intracellular medium appeared to alter the activity and/or accessibility of several enzymes. The metabolism of inositol phosphates appears to occur predominantly in the follicle cells surrounding the oocyte, as collagenase treatment followed by defolliculation greatly diminished the rates of metabolism of several inositol phosphates. The presence in Xenopus ovarian follicles of a 3'-phosphomonoesterase activated by protein kinase C in addition to the well-known 3'-kinase suggests that, by forming a reversible interconversion between Ins(1,4,5)P3 and Ins(1,3,4,5)P4, this tissue may have the potential to prolong stimulatory signals on binding of appropriate agonists to receptors.  相似文献   

6.
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   

7.
Ins(1,4,5)P3 5-phosphatase catalyses the dephosphorylation of Ins(1,4,5)P3 in the 5 position. At 1 microM Ins(1,4,5)P3, 10-15% of total activity of a bovine brain homogenate was measured in the soluble fraction, whereas 85-90% was in the particulate fraction. Particulate activity could be solubilized by cholate or, to a lower extent, by 2 M KCl. Two soluble enzymes (type I and type II) could be fractionated by DEAE-Sephacel chromatography. Soluble activities have been further purified by blue-Sepharose, Sephacryl S-200 and phosphocellulose chromatography. Specific activities reached 10-30 mumol.min-1 mg protein-1 for type I and were 10-20 times lower for type II. Type I and type II Ins(1,4,5)P3 5-phosphatase displayed different Km values and molecular masses, as estimated by gel filtration. Type I dephosphorylated both Ins(1,4,5)P3 and Ins(1,3,4,5)P4; in contrast, type II specifically dephosphorylated Ins(1,4,5)P3 but not Ins(1,3,4,5)P4. Type I Ins(1,4,5)P3 5-phosphatase eluted as a single peak of activity with an apparent molecular mass of 51 kDa when gel filtration was performed in the presence of cholate. This molecular mass is identical to the molecular mass estimated for the particulate Ins(1,4,5)P3 5-phosphatase that was solubilized by cholate. Km values for Ins(1,4,5)P3 and Ins(1,3,4,5)P4 obtained with type I Ins(1,4,5)P3 5-phosphatase were 11 microM and 1 microM, respectively. Similar values were obtained with particulate Ins(1,4,5)P3 5-phosphatase. In conclusion, the catalytic domains of type I and particulate Ins(1,4,5)P3 5-phosphatase activity may be very similar, if not identical, but different from type II phosphatase.  相似文献   

8.
Basal and stimulated levels of inositol phosphates were determined in the protozoan Paramecium labelled with myo-[3H]inositol. Under resting conditions, intracellular InsP6 (phytic acid), InsP5 and InsP4 concentrations were 140, 10 and 2 microM, respectively. InsP5 was comprised of 56% Ins(1,2,3,4,5)P5 and/or Ins(1,2,3,5,6)P5, 40% Ins(1,2,4,5,6)P5 and/or Ins(2,3,4,5,6)P5 and small amounts of Ins(1,3,4,5,6)P5 and Ins(1,2,3,4,6)P5. InsP4 was mainly Ins(1, 4, 5, 6)P4 and/or Ins(3, 4, 5, 6)P4. Other inositol phosphates were not detected at a detection limit of 50-85 nM. Using various depolarizing and hyperpolarizing stimuli, no significant changes in level of inositol phosphates were observed in vivo, indicating that in the ciliate a contribution of inositol phosphates to signal-transduction mechanisms is unlikely. In homogenates prepared from myo-[3H]inositol-labelled cells, a marked relative increase in InsP3 and InsP4 over the concentrations in vivo was observed. These inositol phosphates were identified as degradation products of endogenous InsP6. A novel separation methodology for inositol phosphates was established to allow unequivocal assignment of phosphate locations of all dephosphorylated InsP6-derived products. The dephosphorylation was catalyzed by a phytase-like enzyme with a molecular mass of 240 kDa, most likely of a hexameric structure. The enzyme had a pH optimum of 7.0 and did not require divalent cations for activity. Substrate concentrations above 300 microM were inhibitory. Dephosphorylation of InsP6 by the Paramecium enzyme differs from that of phytases from plants in that it proceeds via a sequential release of phosphate groups from positions 6, 5, 4 and 3 of the myo-inositol ring or/and positions 4, 5, 6 and 1.  相似文献   

9.
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   

10.
InsP6 is an intracellular signal with several proposed functions that is synthesized by IP5K [Ins(1,3,4,5,6)P5 2-kinase]. In the present study, we overexpressed EGFP (enhanced green fluorescent protein)-IP5K fusion proteins in NRK (normal rat kidney), COS7 and H1299 cells. The results indicate that there is spatial microheterogeneity in the intracellular localization of IP5K that could also be confirmed for the endogenous enzyme. This may facilitate changes in InsP6 levels at its sites of action. For example, overexpressed IP5K showed a structured organization within the nucleus. The kinase was preferentially localized in euchromatin and nucleoli, and co-localized with mRNA. In the cytoplasm, the overexpressed IP5K showed locally high concentrations in discrete foci. The latter were attributed to stress granules by using mRNA, PABP [poly(A)-binding protein] and TIAR (TIA-1-related protein) as markers. The incidence of stress granules, in which IP5K remained highly concentrated, was further increased by puromycin treatment. Using FRAP (fluorescence recovery after photobleaching) we established that IP5K was actively transported into the nucleus. By site-directed mutagenesis we identified a nuclear import signal and a peptide segment mediating the nuclear export of IP5K.  相似文献   

11.
The localization of the human erythrocyte membrane Ins(1,3,4,5)P4 3-phosphatase was investigated by saponin permeabilization of resealed 'isoionic' erythrocyte ghosts. This enzyme is active at the inner face of the plasma membrane, at the same site as a specific 5-phosphatase that degrades both Ins (1,4,5)P3 and Ins(1,3,4,5)P4. In the presence of EDTA, Ins(1,4,5)P3 was the only product of Ins(1,3,4,5)P4 metabolism. However, when Mg2+ was present both the 5-phosphatase and the 3-phosphatase attacked Ins (1,3,4,5)P4, directly forming Ins(1,3,4)P3 and Ins(1,4,5)P3;some Ins(1,4)P2 was also formed as a product of 5-phosphatase attack on the liberated Ins(1,4,5)P3. The Ins(1,3,4,5)P4 3-phosphatase was potently activated by KCl, thus making the route of metabolism of Ins(1,3,4,5)P4 by erythrocyte ghosts strikingly sensitive to variations in ionic strength: at 'cytosolic' K+ and Mg2+ levels, 3-phosphatase activity slightly predominated over 5-phosphatase. Ins(1,3,4,5)P4 3-phosphatase was potently inhibited by Ins-(1,3,4,5,6)P5 and InsP6 at levels lower than those often observed within cells. This leaves open the question as to whether the cellular function of inositol polyphosphate 3-phosphatase is to participate in a physiological cycle that interconverts Ins(1,3,4,5)P4 and Ins(1,4,5)P3 or to metabolize other inositol polyphosphates in the cytosol compartment of cells.  相似文献   

12.
The uptake of myo-inositol and its incorporation into the phosphoinositides and inositol phosphates of the phosphatidylinositol (PtdIns) signal transduction system by in vivo preimplantation cattle embryos was investigated using [(3)H] myo-inositol. Uptake of inositol was examined in two-cell and four-cell embryos (day 2 after insemination), morulae (day 6) and early blastocysts (day 7). Uptake in all stages examined was largely sodium-dependent indicating the presence of a sodium-dependent inositol transporter. Uptake of inositol did not vary significantly from two-cell to early blastocyst stages when expressed either on a per embryo or a per microg of protein basis. Incorporation of inositol into the three phosphoinositides, PtdIns, PtdInsP, and PtdInsP(2), was detectable at all stages examined. In contrast, incorporation of inositol into inositol phosphates was not detected until blastocyst formation at day 7. The second messenger, Ins(1,4,5)P(3), was first detected in day 7 blastocysts.  相似文献   

13.
We have identified, isolated, and characterized a second inositol polyphosphate-5-phosphatase enzyme from the soluble fraction of human platelets. The enzyme hydrolyzes inositol 1,4,5-trisphosphate (Ins (1,4,5)P3) to inositol 1,4-bisphosphate (Ins(1,4)P2) with an apparent Km of 24 microM and a Vmax of 25 mumol of Ins(1,4,5)P3 hydrolyzed/min/mg of protein. The enzyme hydrolyzes inositol (1,3,4,5)-tetrakisphosphate (Ins(1,3,4,5)P4) at a rate of 1.3 mumol of Ins(1,3,4,5)P4 hydrolyzed/min/mg of protein with an apparent Km of 7.5 microM. The enzyme also hydrolyzes inositol 1,2-cyclic 4,5-trisphosphate (cIns(1:2,4,5)P3) and Ins(4,5)P2. We purified this enzyme 2,200-fold from human platelets. The enzyme has a molecular mass of 75,000 as determined by both sodium dodecyl sulfate-polyacrylamide gel electrophoresis and by gel filtration chromatography. The enzyme requires magnesium ions for activity and is not inhibited by calcium ions. The 75-kDa inositol polyphosphate-5-phosphatase enzyme differs from the previously identified platelet inositol polyphosphate-5-phosphatase as follows: molecular size (75 kDa versus 45 kDa), affinity for Ins(1,3,4,5)P4 (Km 7.5 microM versus 0.5 microM), Km for Ins(1,4,5)P3 (24 microM versus 7.5 microM), regulation by protein kinase C, wherein the 45-kDa enzyme is phosphorylated and activated while the 75-kDa enzyme is not. The 75-kDa enzyme is inhibited by lower concentrations of phosphate (IC50 2 mM versus 16 mM for the 45-kDa enzyme) and is less inhibited by Ins(1,4)P2 than is the 45-kDa enzyme. The levels of inositol phosphates that act in calcium signalling are likely to be regulated by the interplay of these two enzymes both found in the same cell.  相似文献   

14.
M Lemos  J E Dumont  C Erneux 《FEBS letters》1989,249(2):321-323
Ins(1,4,5)P3 5-phosphatase catalyzes the dephosphorylation of Ins(1,4,5)P3 in the 5-position. In a high speed soluble fraction of bovine brain, there are two soluble 5-phosphatases: type I and type II. The purified Ins(1,4,5)P3 5-phosphatase type I exhibits a major silver-stained band of 43 kDa on denaturing (SDS) gels. It is possible to extract the 5-phosphatase activity form a duplicate lane after gel electrophoresis. The 43 kDa region contains the extractable Ins(1,4,5)P3 5-phosphatase activity.  相似文献   

15.
Inositol 1,3,4,5-tetrakisphosphates (Ins(1,3,4,5)P4), 32P-labelled in positions 4 and 5 were prepared enzymatically, using [4-32P]-phosphatidylinositol 4-phosphate (PtdInsP) and [5-32P]phosphatidylinositol 4,5-bisphosphate (PtdInsP2) as substrates, respectively. Degradation studies of Ins(1,3,4,5)P4, using an enriched phosphatase preparation from porcine brain cytosol, led to the formation of two inositol trisphosphate isomers which were identified as inositol 1,3,4-trisphosphate (Ins(1,3,4)P3) and inositol 1,4,5-trisphosphate (Ins(1,4,5)P3). This novel degradation pathway of Ins(1,3,4,5)P4 to Ins(1,4,5)P3 provides an additional source for the generation of Ins(1,4,5)P3, involving a 3-phosphatase.  相似文献   

16.
The human inositol phosphate multikinase (IPMK, 5-kinase) has a preferred 5-kinase activity over 3-kinase and 6-kinase activities and a substrate preference for inositol 1,3,4,6-tetrakisphosphate (Ins(1,3,4,6)P4) over inositol 1,4,5-trisphosphate (Ins(1,4,5)P3) and inositol 1,3,4,5-tetrakisphosphate (Ins(1,3,4,5)P4). We now report that the recombinant human protein can catalyze the conversion of inositol 1,4,5,6-tetrakisphosphate (Ins(1,4,5,6)P4) to Ins(1,3,4,5,6)P5 in vitro; the reaction product was identified by HPLC to be Ins(1,3,4,5,6)P5. The apparent Vmax was 42 nmol of Ins(1,3,4,5,6)P5 formed/min/mg protein, and the apparent Km was 222 nM using Ins(1,3,4,6)P4 as a substrate; the catalytic efficiency was similar to that for Ins(1,4,5)P3. Stable over-expression of the human protein in HEK-293 cells abrogates the in vivo elevation of Ins(1,4,5,6)P4 from the Salmonella dublin SopB protein. Hence, the human 5-kinase may also regulate the level of Ins(1,4,5,6)P4 and have an effect on chloride channel regulation.  相似文献   

17.
In the rat pancreatoma cell line, AR4-2J, three inositol tetrakisphosphate isomers were identified, (1,3,4,6), (1,3,4,5), (3,4,5,6), which were increased during activation of phospholipase C by bombesin. Two other isomers were identified, (1,4,5,6) and a fifth isomer which was either (1,2,3,4) or (1,2,3,6), which have not previously been detected in any cell type. To study the metabolic interrelationships between these compounds and inositol 1,3,4,5,6-pentakisphosphate in the intact cell, their turnover was assessed under different protocols of [3H]myo-inositol labeling; the inositol phosphates were labeled to near steady state or under conditions where either rapidly or slowly turning over inositol polyphosphates were preferentially labeled. The relative specific radioactivities of inositol 1,4,5-trisphosphate, inositol 1,3,4,5-tetrakisphosphate, inositol 1,3,4-trisphosphate, and inositol 1,3,4,6-tetrakisphosphate were very similar in bombesin-stimulated cells, consistent with the pathway for the conversion of inositol 1,4,5-trisphosphate to the other three inositol polyphosphates. Compared with these inositol phosphates, the turnover of inositol 1,3,4,5,6-pentakisphosphate was slow. An accumulation of radioactivity into inositol 1,3,4,5,6-pentakisphosphate was observed only under labeling conditions where its relative specific radioactivity was substantially below that of inositol 1,3,4,6-tetrakisphosphate. This indicated that the precursor for de novo synthesis of inositol 1,3,4,5,6-pentakisphosphate was inositol 1,3,4,6-tetrakisphosphate. Bombesin stimulated the net breakdown of inositol 1,3,4,5,6-pentakisphosphate and increased the level of inositol 3,4,5,6-tetrakisphosphate; the relative specific radioactivities of these two compounds were similar under all conditions. These data led to the novel proposal that inositol 3,4,5,6-tetrakisphosphate is the product of inositol 1,3,4,5,6-pentakisphosphate breakdown. This reaction was apparently stimulated by a regulated change in the enzyme(s) which interconvert inositol 1,3,4,5,6-pentakisphosphate and inositol 3,4,5,6-tetrakisphosphate.  相似文献   

18.
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   

19.
The inositol phosphates are ubiquitous metabolites in eukaryotes, of which the most abundant are inositol hexakisphosphate (InsP 6) and inositol 1,3,4,5,6-pentakisphosphate [Ins(1,3,4,5,6)P 5)]. These two compounds, poorly understood functionally, have complicated complexation and solid formation behaviours with multivalent cations. For InsP 6, we have previously described this chemistry and its biological implications (Veiga et al. in J Inorg Biochem 100:1800, 2006; Torres et al. in J Inorg Biochem 99:828, 2005). We now cover similar ground for Ins(1,3,4,5,6)P 5, describing its interactions in solution with Na+, K+, Mg2+, Ca2+, Cu2+, Fe2+ and Fe3+, and its solid-formation equilibria with Ca2+ and Mg2+. Ins(1,3,4,5,6)P 5 forms soluble complexes of 1:1 stoichiometry with all multivalent cations studied. The affinity for Fe3+ is similar to that of InsP 6 and inositol 1,2,3-trisphosphate, indicating that the 1,2,3-trisphosphate motif, which Ins(1,3,4,5,6)P 5 lacks, is not absolutely necessary for high-affinity Fe3+ complexation by inositol phosphates, even if it is necessary for their prevention of the Fenton reaction. With excess Ca2+ and Mg2+, Ins(1,3,4,5,6)P 5 also forms the polymetallic complexes [M4(H2L)] [where L is fully deprotonated Ins(1,3,4,5,6)P 5]. However, unlike InsP 6, Ins(1,3,4,5,6)P 5 is predicted not to be fully associated with Mg2+ under simulated cytosolic/nuclear conditions. The neutral Mg2+ and Ca2+ complexes have significant windows of solubility, but they precipitate as [Mg4(H2L)]·23H2O or [Ca4(H2L)]·16H2O whenever they exceed 135 and 56 μM in concentration, respectively. Nonetheless, the low stability of the [M4(H2L)] complexes means that the 1:1 species contribute to the overall solubility of Ins(1,3,4,5,6)P 5 even under significant Mg2+ or Ca2+ excesses. We summarize the solubility behaviour of Ins(1,3,4,5,6)P 5 in straightforward plots.  相似文献   

20.
Appropriate pollen germination is crucial for plant reproduction. Previous studies have revealed the importance of dehydration in maintaining pollen dormancy; here, we show that phosphatidylinositol pathway-controlled Ins(1,4,5)P(3)/Ca(2+) levels are crucial for maintaining pollen dormancy in Arabidopsis thaliana. An interesting phenotype, precocious pollen germination within anthers, results from a disruption of inositol polyphosphate 5-phosphatase 12 (5PT12). The knockout mutant 5pt12 has normal early pollen development and pollen dehydration, and exhibits hypersensitive ABA responses, indicating that precocious pollen germination is not caused either by abnormal dehydration or by suppressed ABA signaling. Deficiency of 5PT13 (a close paralog of 5PT12) synergistically enhances precocious pollen germination. Both basal Ins(1,4,5)P(3) levels and endogenous Ca(2+) levels are elevated in pollen from 5pt12 mutants, and 5pt12 5pt13 double mutants show an even higher precocious germination rate along with much higher levels of Ins(1,4,5)P(3)/Ca(2+). Strikingly, exogenous Ca(2+) stimulates the germination of wild-type pollen at floral stage 12, even in very low humidity, both in vitro and in vivo, and treatment with BAPTA, a [Ca(2+)](cyt) inhibitor, reduces the precocious pollen germination rates of 5pt12, 5pt13 and 5pt12 5pt13 mutants. These results indicate that the increase in the levels of Ins(1,4,5)P(3)/Ca(2+) caused by deficiency of inositol polyphosphate 5-phosphatases is sufficient to break pollen dormancy and to trigger early germination. The study reveals that independent of dehydration, the control of Ins(1,4,5)P(3)/Ca(2+) levels by Inositol polyphosphate 5-phosphatases is crucial for maintaining pollen dormancy.  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