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.
When screening an Escherichia coli gene library for myo-inositol hexakisphosphate (InsP6) phosphatases (phytases), we discovered that the agp-encoded acid glucose-1-phosphatase also possesses this activity. Purified Agp hydrolyzes glucose-1-phosphate, p-nitrophenyl phosphate, and InsP6 with pH optima, 6.5, 3.5, and 4.5, respectively, and was stable when incubated at pH values ranging from 3 to 10. Glucose-1-phosphate was hydrolyzed most efficiently at 55 degrees C. while InsP6 and p-nitrophenyl phosphate were hydrolyzed maximally at 60 degrees C. The Agp exhibited Km values of (0.39 mM, 13 mM, and 0.54 mM for the hydrolysis of glucose-1-phosphate, p-nitrophenyl phosphate, and InsP6, respectively. High-pressure liquid chromatography (HPLC) analysis of inositol phosphate hydrolysis products of Agp demonstrated that the enzyme catalyzes the hydrolysis of phosphate from each of InsP6, D-Ins(1,2,3,4,5)P5, Ins(1,3,4,5,6)P5, and Ins(1,2,3,4,6)P5, producing D/L-Ins(1,2,4,5,6)P5. D-Ins(1,2,4,5)P4, D/L-Ins(1,4,5,6)P4 and D/L-Ins(1,2,4,6)P4, respectively. These data support the contention that Agp is a 3-phosphatase.  相似文献   

2.
Using a combination of high-performance ion chromatography analysis and kinetic studies, the stereospecificity of myo-inositol hexakisphosphate dephosphorylation by the phytate-degrading enzyme P2 of Escherichia coli was established. High-performance ion chromatography revealed that the phytate-degrading enzyme P2 of E. coli degrades myo-inositol hexakisphosphate by stepwise dephosphorylation via D/L-Ins(1,2,3,4,5)P(5), D/L-Ins(2,3,4,5)P(4), D/L-Ins(2,4,5)P(3) or D/L-Ins(1,2,4)P(3), D/L-Ins(1,2)P(2) or Ins(2, 5)P(2) or D/L-Ins(4,5)P(2) to finally Ins(2)P or Ins(5)P. Kinetic parameters for myo-inositol pentakisphosphate hydrolysis by E. coli and wheat phytase, respectively, showed that the myo-inositol pentakisphosphate intermediate produced either by the phytate-degrading enzyme of wheat or E. coli are not identical. The absolute configuration of the myo-inositol pentakisphosphate isomer produced by the E. coli enzyme was determined by taking into consideration that wheat phytase produces predominantly the D-Ins(1, 2,3,5,6)P(5) isomer (Lim, P.E., Tate, M.E., 1973. The phytases: II. Properties of phytase fraction F(1) and F(2) from wheat bran and the myo-inositol phosphates produced by fraction F(2). Biochim. Biophys. Acta 302, 326-328). The data demonstrate that the phytate-degrading enzyme P2 of E. coli dephosphorylates myo-inositol hexakisphosphate in a stereospecific way by sequential removal of phosphate groups via D-Ins(1,2,3,4,5)P(5), D-Ins(2,3,4,5)P(4), D-Ins(2,4,5)P(3), Ins(2,5)P(2) to finally Ins(2)P (notation 6/1/3/4/5).  相似文献   

3.
In cerebral cortex of rats treated with increasing doses of LiCl, the relative concentrations of Ins(1)P, Ins(4)P and Ins(5)P (when InsP is a myo-inositol phosphate) are approx. 10:1:0.2 at all doses. In rats treated with LiCl followed by increasing doses of pilocarpine a similar relationship occurs. myo-Inositol-1-phosphatase (InsP1ase) from bovine brain hydrolyses Ins(1)P, Ins(4)P and Ins(5)P at comparable rates, and these substrates have similar Km values. The hydrolysis of Ins(4)P is inhibited by Li+ to a greater degree than is hydrolysis of Ins(1)P and Ins(5)P. D-Ins(1,4,5)P3 and D-Ins(1,4)P2 are neither substrates nor inhibitors of InsP1ase. A dialysed high-speed supernatant of rat brain showed a greater rate of hydrolysis of Ins(1)P than of D-Ins(1,4)P2 and a lower sensitivity of the bisphosphate hydrolysis to LiCl, as compared with the monophosphate. That enzyme preparation produced Ins(4)P at a greater rate than Ins(1)P when D-Ins(1,4)P2 was the substrate. The amount of D-Ins(3)P [i.e. L-Ins(1)P, possibly from D-Ins(1,3,4)P3] is only 11% of that of D-Ins(1)P on stimulation with pilocarpine in the presence of Li+. DL-Ins(1,4)P2 was hydrolysed by InsP1ase to the extent of about 50%; both Ins(4)P and Ins(1)P are products, the former being produced more rapidly than the latter; apparently L-Ins(1,4)P2 is a substrate for InsP1ase. Li+, but not Ins(2)P, inhibited the hydrolysis of L-Ins(1,4)P2. The following were neither substrates nor inhibitors of InsP1ase; Ins(1,6)P2, Ins(1,2)P2, Ins(1,2,5,6)P4, Ins(1,2,4,5,6)P5, Ins(1,3,4,5,6)P5 and phytic acid. myo-Inositol 1,2-cyclic phosphate was neither substrate nor inhibitor of InsP1ase. We conclude that the 10-fold greater tissue contents of Ins(1)P relative to Ins(4)P in both stimulated and non-stimulated rat brain in vivo are the consequence of a much larger amount of PtdIns metabolism than polyphosphoinositide metabolism under these conditions.  相似文献   

4.
Brearley CA  Hanke DE 《Plant physiology》2000,122(4):1209-1216
Using a permeabilization strategy to introduce Ins(3,4,5,6) P(4) into mesophyll protoplasts of Commelina communis, we have identified Ins(3,4,5,6) P(4) 1-kinase activity in mesophyll cells. Multiple InsP(3) isomers were identified in Spirodela polyrhiza and Arabidopsis. Only two of these, Ins(1,2,3) P(3) and Ins(3,4,6) P(3), have previously been identified in plants and only in monocots. The isomers detected in S. polyrhiza included D- and/or L-Ins(3,4,5) P(3), D- and/or L-Ins(3,5,6) P(3), and D- and/or L-Ins(2,4,5) P(3). Ins(1,4,5) P(3), if present, was only a tiny fraction of total InsP(3) species. We have also identified inositol polyphosphate phosphatase activities, Ins(3,4,5,6) P(4) 6-phosphatase and Ins(3,4, 5, 6) P(4) 4-phosphatase, whose action on endogenous inositol polyphosphates explains the presence of D- and/or L-Ins(3,4,5) P(3) and D- and/or L-Ins(3,5,6) P(3) in mesophyll cells. Inositol trisphosphates identified in Arabidopsis include Ins(1,2,3) P(3) and D- and/or L-Ins(3,4,6) P(3), suggesting that dicots may share pathways of InsP(6) biosynthesis and breakdown in common with monocots.  相似文献   

5.
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   

6.
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   

7.
Thrombin-stimulated (10 s) human platelets produce Ins(1,4,5)P3 and an additional inositol trisphosphate (InsP3), in approximately a 1:20 ratio. The major InsP3 co-migrates with Ins(1,3,4)P3 on strong-anion-exchange h.p.l.c. To identify this species unequivocally, we treated putative Ins(1,3,4)P3 obtained from thrombin-stimulated myo-[3H]inositol-labelled platelets with NaIO4/NaBH4 or 4-phosphomonoesterase. The products indicate that the major InsP3 is at least 90% D-Ins(1,3,4)P3. D-[3H]Ins(1,3,4)P3 added to saponin-permeabilized platelets is hydrolysed to an InsP2 (7.8%) and phosphorylated by a kinase to yield an inositol polyphosphate (0.9%) in 5 min. The phosphorylation product co-migrates with Ins(1,3,4,6)P4 on Partisphere WAX h.p.l.c. Under similar conditions, L-[3H]Ins(1,3,4)P3 is dephosphorylated but not phosphorylated. Relative phosphatase:kinase ratios are 8.7:1 (Vmax. values) and 0.86:1 (Km values) with respect to D-Ins(1,3,4)P3. The kinase activity is predominantly cytosolic (96.8% of total activity) in freeze-thaw-disrupted platelets, and the accumulation of its product is Ca2(+)-dependent. The activity is identified as a 6-kinase on the basis of its product's insensitivity to 5-phosphomonoesterase, resistance to periodate oxidation and co-migration with standard Ins(1,3,4,6)P4 on h.p.l.c. Incubation of platelets with beta-phorbol dibutyrate (beta-PDBu, 76 nM), causing activation of protein kinase C, results in a 57.5% inhibition (reversible by the protein kinase C inhibitor staurosporine) of Ins(1,3,4,6)P4 accumulation. alpha-PDBu, which does not stimulate protein kinase C, has no effect. Stimulation of intact platelets with thrombin results in the production of Ins(1,3,4,6)P4 (1.4-fold rise in 30 s) and Ins(1,3,4,5)P4, with the latter being the major InsP4 species. Accumulation of Ins(1,3,4,6)P4 is slightly delayed in comparison with Ins(1,3,4)P3 and is relatively small. We propose that the major route of Ins(1,3,4)P3 metabolism in stimulated human platelets is via phosphatase action.  相似文献   

8.
Inositol polyphosphatases (IPPases) play an important role in the metabolism of inositol polyphosphates, a class of molecules involved in signal transduction. Here we characterize 2 new protein tyrosine phosphatase-like IPPases (PhyAsl and PhyBsl) cloned from Selenomonas lacticifex that can hydrolyze myo-inositol hexakisphosphate (InsP6) in vitro. To determine their preferred substrates and stereospecificity of InsP6 dephosphorylation, a combination of kinetic and high-performance ion pair chromatography studies were conducted. Despite only 33% amino acid sequence identity between them, both enzymes display strict specificity for IPP substrates and cleave InsP6 primarily at the d-3-phosphate position (>90%). Furthermore, both enzymes predominantly degrade InsP6 to Ins(2)P via identical and very specific routes of dephosphorylation (3,4,5,6,1). Despite these similarities, PhylAsl is shown to have a slight kinetic preference for the major inositol pentakisphosphate intermediate in its InsP6 hydrolysis pathway, whereas PhyBsl displays a unique and substantial preference for an inositol tetrakisphosphate intermediate.  相似文献   

9.
10.
L1210 lymphoma cells were permeabilized with digitonin, and the ability of Ins(2,4,5)P3 and Ins(1,3,4,5)P4 to mobilize intracellular Ca2+ was studied. At high doses of Ins(2,4,5)P3 Ca2+ was rapidly released from intracellular stores, and prior or subsequent addition of Ins(1,3,4,5)P4 had no discernible effect. However, the Ca2(+)-mobilizing action of low (threshold or just above) concentrations of Ins(2,4,5)P3 was markedly enhanced by Ins(1,3,4,5)P4, which alone caused no mobilization of Ca2+; this phenomenon was shown not to be due to protection of Ins(2,4,5)P3 by the Ins(1,3,4,5)P4 against hydrolysis. The ability of the pre-addition of Ins(1,3,4,5)P4 to enhance subsequent Ins(2,4,5)P3-induced Ca2+ mobilization was always seen whether or not the free Ca2+ concentration was low (pCa = 7) or high (pCa = 6). However, at low Ca2+, Ins(1,3,4,5)P4 could cause a further mobilization if added after the Ins(2,4,5)P3, whereas at higher Ca2+ values Ins(1,3,4,5)P4 was only able to affect Ca2+ if added before Ins(2,4,5)P3. These effects of Ins(1,3,4,5)P4 were not, at the same concentration, mimicked by a random mixture of InsP4 isomers obtained by partial acid hydrolysis of phytic acid, by Ins(1,3,4)P3 or by Ins(1,3,4,5,6)P5, and they were shown not to be due to enzymic generation of Ins(1,4,5)P3 from Ins(1,3,4,5)P4 by (a) the absence of any detectable production of Ins(1,4,5)P3 if radiolabelled Ins(1,3,4,5)P4 was used, or (b) the observation that Ins(1,3,4,5,6)P5 could mimic Ins(1,3,4,5)P4 provided that higher doses were used; this inositol phosphate, when added radiolabelled, yielded only trace quantities of D/L-Ins(1,4,5,6)P4, which itself does not mobilize Ca2+. We interpret these results overall to mean that in these cells there is a small proportion of the Ins(2,4,5)P3-mobilizable Ca2+ pools which can only be mobilized in the presence of Ins(1,3,4,5)P4 [or at the least, Ins(1,3,4,5)P4 can help Ins(2,4,5)P3 to gain access to them]. The significance of this conclusion is discussed in the light of current concepts of the second messenger function of Ins(1,3,4,5)P4.  相似文献   

11.
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   

12.
We studied the dephosphorylation pathway for Ins(1,3,4)P3 (inositol 1,3,4-trisphosphate) by liver homogenates and soluble and particulate subfractions incubated in media resembling physiological ionic strength and pH. Ins(1,3,4)P3 was dephosphorylated to two InsP2 (inositol bisphosphate) isomers, one of which is Ins(3,4)P2 [Shears, Parry, Tang, Irvine, Michell & Kirk (1987) Biochem. J. 246, 139-147]. The second InsP2 is the 1,3 isomer. Ins(3,4)P2 is dephosphorylated to inositol 3-phosphate by an enzyme activity located in both soluble and particulate fractions. The phosphatase(s) that attacks Ins(1,3)P2 was largely soluble, but we have not determined which phosphate(s) is removed. When the initial substrate concentration was 1 nM, the rate of dephosphorylation of Ins(1,4)P2 greater than Ins(1,3)P2 greater than Ins(3,4)P2. None of these bisphosphates was phosphorylated when incubated with liver homogenates and 5 mM-ATP, but their rates of dephosphorylation were then decreased.  相似文献   

13.
Using a combination of High-Performance Ion Chromatography analysis and kinetic studies, the pathway of myo-inositol hexakisphosphate dephosphorylation by a phytase from a Malaysian waste-water bacterium was established. The data demonstrate that the phytase preferably dephosphorylates myo-inositol hexakisphosphate in a stereospecific way by sequential removal of phosphate groups via D-I(1,2,3,4,5)P5, D-I(2,3,4,5)P4, D-I(2,3,4)P3, D-I(2,3)P2 to finally I(2)P. It was estimated that more than 90% of phytate hydrolysis occurs via D-I(1,2,3,4,5)P5. Thus, the phytase from the Malaysian waste-water bacterium has to be considered a 6-phytase (E.C. 3.1.3.26). A second pathway of minor importance could be proposed which is in accordance with the results obtained from analysis of the dephosphorylation products formed by the action of the phytase under investigation on myo-inositol hexakisphosphate. It proceeds via D/L-I(1,2,4,5,6)P5, D/L-I(1,2,4,5)P4, D/L-I(1,2,4)P3, D/L-I(2,4)P2 to finally I(2)P.  相似文献   

14.
1. We have studied the metabolism of Ins(1,3,4,5)P4 (inositol 1,3,4,5-tetrakisphosphate) by rat liver homogenates incubated in a medium resembling intracellular ionic strength and pH. 2. Ins(1,3,4,5)P4 was dephosphorylated to a single inositol trisphosphate product, Ins(1,3,4)P3 (inositol 1,3,4-trisphosphate), the identity of which was confirmed by periodate degradation, followed by reduction and dephosphorylation to yield altritol. 3. The major InsP2 (inositol bisphosphate) product was inositol 3,4-bisphosphate [Shears, Storey, Morris, Cubitt, Parry, Michell & Kirk (1987) Biochem. J. 242, 393-402]. Small quantities of a second InsP2 product was also detected in some experiments, but its isomeric configuration was not identified. 4. The Ins(1,3,4,5)P4 5-phosphatase activity was primarily associated with plasma membranes. 5. ATP (5 mM) decreased the membrane-associated Ins(1,4,5)P3 5-phosphatase and Ins(1,3,4,5)P4 5-phosphatase activities by 40-50%. This inhibition was imitated by AMP, adenosine 5'-[beta gamma-imido]triphosphate, adenosine 5'-[gamma-thio]triphosphate or PPi, but not by adenosine or Pi. A decrease in [ATP] from 7 to 3 mM halved the inhibition of Ins(1,3,4,5)P4 5-phosphatase activity, but the extent of inhibition was not further decreased unless [ATP] less than 0.1 mM. 6. Ins(1,3,4,5)P4 5-phosphatase was insensitive to 50 mM-Li+, but was inhibited by 5 mM-2,3-bisphosphoglycerate. 7. The Ins(1,3,4,5)P4 5-phosphatase activity was unchanged by cyclic AMP, GTP, guanosine 5'-[beta gamma-imido]triphosphate or guanosine 5'-[gamma-thio]triphosphate, or by increasing [Ca2+] from 0.1 to 1 microM. 8. Ins(1,3,4)P3 was phosphorylated in an ATP-dependent manner to an isomer of InsP4 that was partially separable on h.p.l.c. from Ins(1,3,4,5)P4. The novel InsP4 appears to be Ins(1,3,4,6)P4. Its metabolic fate and function are not known.  相似文献   

15.
Inositol 1,4,5-trisphosphate receptors (InsP(3)R) play a key role in intracellular calcium (Ca(2+)) signaling. Three InsP(3)R isoforms are expressed in mammals. Type 1 InsP(3)R (InsP(3)R1) is a predominant neuronal isoform. Neuronal InsP(3)R1 is one of the major substrates of protein kinase A (PKA) phosphorylation. In our previous study (Tang, T. S., Tu, H., Wang, Z., and Bezprozvanny, I. (2003) J. Neurosci. 23, 403-415) we discovered a direct association between InsP(3)R1 and protein phosphatase 1 alpha (PP1 alpha). In functional experiments we demonstrated that phosphorylation by PKA activates InsP(3)R1 and that dephosphorylation by PP1 alpha inhibits InsP(3)R1. To extend these findings, here we investigated the possibility of InsP(3)R1-PKA association. In a series of biochemical experiments we demonstrate the following findings. 1) InsP(3)R1 and PKA associate in the brain. 2) InsP(3)R1-PKA association is mediated by the AKAP9 (Yotiao) multi-functional PKA anchoring protein. 3) InsP(3)R1-AKAP9 association is mediated via the leucine/isoleucine zipper (LIZ) motif in the InsP(3)R1 coupling domain and the fourth LIZ motif in AKAP9. 4) The InsP(3)R association with AKAP9 is specific for type 1 InsP(3)R. 5) Both the SII(+) and the SII(-) coupling domain splice variants of InsP(3)R1 bind to AKAP9. 6) Binding to AKAP9 promotes association of neuronal InsP(3)R1 with the NR1 NMDA receptor; and 7) neuronal InsP(3)R1 associate with PP1 directly via carboxy-terminus and indirectly via AKAP9. The obtained results advance our understanding of cross-talk between cAMP and InsP(3)/Ca(2+) signaling pathways in the brain.  相似文献   

16.
In human HeLa carcinoma cells, histamine causes a dose-dependent formation of inositol phosphates, production of diacylglycerol and a transient rise in intracellular [Ca2+]. These responses are completely blocked by the H1-receptor antagonist pyrilamine. In streptolysin-O-permeabilized cells, formation of inositol phosphates by histamine is strongly potentiated by guanosine 5'-[gamma-thio]triphosphate and inhibited by guanosine 5'-[beta-thio]diphosphate, suggesting the involvement of a GTP-binding protein. Histamine stimulates the rapid but transient formation of Ins(1,4,5)P3, Ins(1,3,4)P3 and InsP4. InsP accumulates in a much more persistent manner, lasting for at least 30 min. Studies with streptolysin-O-permeabilized cells indicate that InsP accumulation results from dephosphorylation of Ins(1,4,5)P3, rather than direct hydrolysis of PtdIns. The rise in intracellular [Ca2+] is biphasic, with a very fast release of Ca2+ from intracellular stores, that parallels the Ins(1,4,5)P3 time course, followed by a more prolonged phase of Ca2+ influx. In individual cells, histamine causes a rapid initial hyperpolarization of the plasma membrane, which can be mimicked by microinjected Ins(1,4,5)P3. Histamine-induced hyperpolarization is followed by long-lasting oscillations in membrane potential, apparently owing to periodic activation of Ca2+-dependent K+ channels. These membrane-potential oscillations can be mimicked by microinjection of guanosine 5'-[gamma-thio]triphosphate, but are not observed after microinjection of Ins(1,4,5)P3. We conclude that H1-receptors in HeLa cells activate a PtdInsP2-specific phospholipase C through participation of a specific G-protein, resulting in long-lasting oscillations of cytoplasmic free Ca2+.  相似文献   

17.
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+]i 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   

18.
Polyphosphoinositide-specific phospholipases (PICs) of the delta-subfamily are ubiquitous in eukaryotes, but an inability to control these enzymes physiologically has been a major obstacle to understanding their cellular function(s). Plc1p is similar to metazoan delta-PICs and is the only PIC in Saccharomyces cerevisiae. Genetic studies have implicated Plc1p in several cell functions, both nuclear and cytoplasmic. Here we show that a brief hypo-osmotic episode provokes rapid Plc1p-catalyzed hydrolysis of PtdIns(4,5)P2 in intact yeast by a mechanism independent of extracellular Ca2+. Much of this PtdIns(4,5)P2 hydrolysis occurs at the plasma membrane. The hydrolyzed PtdIns(4,5)P2 is mainly derived from PtdIns4P made by the PtdIns 4-kinase Stt4p. PtdIns(4,5)P2 hydrolysis occurs normally in mutants lacking Arg82p or Ipk1p, but they accumulate no InsP6, showing that these enzymes normally convert the liberated Ins(1,4,5)P3 rapidly and quantitatively to InsP6. We conclude that hypo-osmotic stress activates Plc1p-catalyzed PtdIns(4,5)P2 at the yeast plasma membrane and the liberated Ins(1,4,5)P3 is speedily converted to InsP6. This ability routinely to activate Plc1p-catalyzed PtdIns(4,5)P2 hydrolysis in vivo opens up new opportunities for molecular and genetic scrutiny of the regulation and functions of phosphoinositidases C of the delta-subfamily.  相似文献   

19.
The conformation and dynamics of the deoxyribose rings of a (nogalamycin)2-d(5'-GCATGC)2 complex have been determined from an analysis of 1H-1H vicinal coupling constants and sums of coupling constants (J1'-2',J1'-2",epsilon 1', epsilon 2' and epsilon 2") measured from one-dimensional n.m.r. spectra and from H-1'-H-2' and H-1'-H-2" cross-peaks in high-resolution phase-sensitive two-dimensional correlation spectroscopy (COSY) and double-quantum-filtered correlation spectroscopy (DQF-COSY) experiments. The value of J3'-4' has also been estimated from the magnitude of H-3'-H-4' cross-peaks in DQF-COSY spectra and H-1'-H-4' coherence transfer cross-peaks in two-dimensional homonuclear Hartman-Hahn spectroscopy (HOHAHA) spectra. The data were analysed, in terms of a dynamic equilibrium between North (C-3'-endo) and South (C-2'-endo) conformers, by using the graphical-analysis methods described by Rinkel & Altona [(1987) J. Biomol. Struct. Dyn. 4,621-649]. The data reveal that the sugars of the 2C-5G and 3A-4T base-pairs, which form the drug-intercalation site, have strikingly different properties. The deoxyribose rings of the 2C-5G base-pair are best described in terms of an equilibrium heavily weighted in favour of the C-2'-endo geometry (greater than 95% 'S'), with a phase angle, P, lying in the range 170-175 degrees and amplitude of pucker between 35 and 40 degrees, as typically found for B-DNA. For the deoxyribose rings of the 3A-4T base-pair, however, the analysis shows that, for 3A, the C-2'-endo and C3'-endo conformers are equally populated, whereas a more limited data set for the 4T nucleotide restricts the equilibrium to within 65-75% C-2'-endo. The deoxyribose rings of the 1G-6C base-pair have populations of 70-80% C-2'-endo, typical of nucleotides at the ends of a duplex. Although drug-base-pair stacking interactions are an important determinant of the enhanced duplex stability of the complex [Searle, Hall, Denny, & Wakelin (1988) Biochemistry 27, 4340-4349], the current findings make it clear that the same interactions can be associated with considerable variations in the degree of local structural dynamics at the level of the sugar puckers.  相似文献   

20.
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   

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