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.
The agonist-dependent hydrolysis of inositol phospholipids was investigated by studying the breakdown of prelabelled lipid or by measuring the accumulation of inositol phosphates. Stimulation of insect salivary glands with 5-hydroxytryptamine for 6 min provoked a rapid disappearance of [3H]phosphatidylinositol 4,5-bisphosphate [PtdIns(4,5)P2] and [3H]phosphatidylinositol 4-phosphate (PtdIns4P) but had no effect on the level of [3H]phosphatidylinositol (PtdIns). The breakdown of PtdIns(4,5)P2 was associated with a very rapid release of inositol 1,4,5-trisphosphate [Ins(1,4,5)P3], which reached a peak 5 1/2 times that of the resting level after 5 s of stimulation. This high level was not maintained but declined to a lower level, perhaps reflecting the disappearance of PtdIns(4,5)P2. 5-Hydroxytryptamine also induced a rapid and massive accumulation of inositol 1,4-bisphosphate [Ins(1,4)P2]. The fact that these increases in Ins(1,4,5)P3 and Ins(1,4)P2 precede in time any increase in the level of inositol 1-phosphate or inositol provides a clear indication that the primary action of 5-hydroxytryptamine is to stimulate the hydrolysis of PtdIns(4,5)P2 to yield diacylglycerol and Ins(1,4,5)P3. The latter is then hydrolysed by a series of phosphomonoesterases to produce Ins(1,4)P2, Ins1P and finally inositol. The very rapid agonist-dependent increases in Ins(1,4,5)P3 and Ins(1,4)P2 suggests that they could function as second messengers, perhaps to control the release of calcium from internal pools. The PtdIns(4,5)P2 that is used by the receptor mechanism represents a small hormone-sensitive pool that must be constantly replenished by phosphorylation of PtdIns. Small changes in the size of this small energy-dependent pool of polyphosphoinositide will alter the effectiveness of the receptor mechanism and could account for phenomena such as desensitization and super-sensitivity.  相似文献   

2.
We have synthesized two photolabile arylazido-analogues of Ins(1,4,5)P3 selectively substituted at the 1-phosphate group for determination of Ins(1,4,5)P3-binding proteins. These two photoaffinity derivatives, namely N-(4-azidobenzoyl)aminoethanol-1-phospho-D-myo-inositol 4,5-bisphosphate (AbaIP3) and N-(4-azidosalicyl)aminoethanol-1-phospho-D-myo-inositol 4,5-bisphosphate (AsaIP3), bind to high affinity Ins(1,4,5)P3-specific binding sites at a 9-fold lower affinity (Kd = 66 and 70 nM) than Ins(1,4,5)P3 (Kd = 7.15 nM) in a fraction from rat pancreatic acinar cells enriched in endoplasmic reticulum (ER). Other inositol phosphates tested showed comparable (DL-myo-inositol 1,4,5-trisphosphothioate, Kd = 81 nM) or much lower affinities for the binding sites [Ins(1,3,4,5)P4, Kd = 4 microM; Ins(1,4)P2, Kd = 80 microM]. Binding of AbaIP3 was also tested on a microsomal preparation of rat cerebellum [Kd = 300 nM as compared with Ins(1,4,5)P3, Kd = 45 nM]. Ca2+ release activity of the inositol derivatives was tested with AbaIP3. It induced a rapid and concentration-dependent Ca2+ release from the ER fraction [EC50 (dose producing half-maximal effect) = 3.1 microM] being only 10-fold less potent than Ins(1,4,5)P3 (EC50 = 0.3 microM). From the two radioactive labelled analogues ([3H]AbaIP3 and 125I-AsIP3) synthesized, the radioiodinated derivative was used for photoaffinity labelling. It specifically labelled three proteins with apparent molecular masses of 49, 37 and 31 kDa in the ER-enriched fraction. By subfractionation of this ER-enriched fraction on a Percoll gradient the 37 kDa Ins(1,4,5)P3 binding protein was obtained in a membrane fraction which showed the highest effect in Ins(1,4,5)P3-inducible Ca2+ release (fraction P1). The other two Ins(1,4,5)P3-binding proteins, of 49 and 31 kDa, were obtained in fraction P2, in which Ins(1,4,5)P3-induced Ca2+ release was half of that obtained in fraction P1. We conclude from these data that the 37 kDa and/or the 49 and 31 kDa proteins are involved in Ins(1,4,5)P3-induced Ca2+ release from the ER of rat pancreatic acinar cells.  相似文献   

3.
The effects of inositol 1,4,5-trisphosphate (Ins(1,4,5)P3), which has been hypothesized to be a chemical transmitter in excitation-contraction coupling in skeletal muscle, on aldolase bound to isolated triad junctions were investigated. Fructose-1,6-bisphosphate aldolase was identified as the major specific binding protein for the Ins(1,4,5)P3 analogue glycolaldehyde (2)-1-phospho-D-myo-inositol 4,5-bisphosphate which can form covalent bonds with protein amino groups by reduction of the Schiff's base intermediate with [3H]NaCNBH3. This analogue, Ins(1,4,5) P3, and the inositol polyphosphates inositol 1,3,4,5-tetrakisphosphate and inositol 1,4-bisphosphate were nearly equipotent in selectively releasing membrane bound aldolase with a K0.5 of about 3 microM. The rank order of the K0.5 values was identical to the KI values for inhibition of aldolase. Aldolase was also released by its substrate fructose 1,6-bisphosphate and by 2,3-bisphosphoglycerate. Ins(1,4,5)P3-induced aldolase release did not disrupt the triad junction; glyceraldehyde-3-phosphate dehydrogenase, a known junctional constituent, was displaced only at much higher Ins(1,4,5)P3 concentrations. Ins(1,4,5)P3 was as effective as fructose 1,6-bisphosphate in releasing aldolase from myofibrils. A finite number of binding sites for aldolase exist on triads (Bmax = 43-47 pmol of tetrameric aldolase exist on triads (Bmax = 43-47 pmol of tetrameric aldolase/mg of triad protein, KD = 23 nM). The junctional foot protein was implicated as an aldolase binding site by affinity chromatography with the junctional foot protein immobilized on Sepharose 4B. The potential consequences of aldolase being bound in the gap between the terminal cisternae and the transverse tubule to inositol polyphosphate and glycolytic metabolism in that local region are discussed.  相似文献   

4.
Inositol-5-phosphatases are important enzymes involved in the regulation of diverse cellular processes from synaptic vesicle recycling to insulin signaling. We describe a comparative study of two representative inositol-5-phosphatases, Schizosaccharomyces pombe synaptojanin (SPsynaptojanin) and human SH2 domain-containing inositol-5-phosphatase SHIP2. We show that in addition to Mg2+, transition metals such as Mn2+, Co2+, and Ni2+ are also effective activators of SPsynaptojanin. In contrast, Ca2+ and Cu2+ are inhibitory. We provide evidence that Mg2+ binds the same site occupied by Ca2+ observed in the crystal structure of SPsynaptojanin complexed with inositol 1,4-bisphosphate (Ins(1,4)P2). Ionizations important for substrate binding and catalysis are defined for the SPsynaptojanin-catalyzed Ins(1,4,5)P3 reaction. Kinetic analysis with four phosphatidylinositol lipids bearing a 5-phosphate and 54 water-soluble inositol phosphates reveals that SP-synaptojanin and SHIP2 possess much broader substrate specificity than previously appreciated. The rank order for SPsynaptojanin is Ins(2,4,5)P3 > phosphatidylinositol-4,5-bisphosphate (PtdIns(4,5)P2) approximately Ins(4,5)P2 approximately Ins(1,4,5)P3 approximately Ins(4,5,6)P3 > PtdIns(3,5)P2 approximately PtdIns(3,4,5)P3 approximately Ins(1,2,4,5)P4 approximately Ins(1,3,4,5)P4 approximately Ins-(2,4,5,6)P4 approximately Ins(1,2,4,5,6)P5. The rank order for SHIP2 is Ins(1,2,3,4,5)P5 > Ins(1,3,4,5)P4 > PtdIns(3,4,5)P4 approximately PtdIns(3,5)P2 approximately Ins(1,4,5,6)P4 approximately Ins(2,4,5,6)P4. Because inositol phosphate isomers elicit different biological activities, the extended substrate specificity for SPsynaptojanin and SHIP2 suggest that these enzymes likely have multiple roles in cell signaling and may regulate distinct pathways. The unique substrate specificity profiles and the importance of 2-position phosphate in binding also have important implications for the design of potent and selective SPsynaptojanin and SHIP2 inhibitors for pharmacological investigation.  相似文献   

5.
1. Phosphatidylinositol 4-phosphate (PtdIns4P) is degraded by isolated membranes from Xenopus laevis oocytes. 2. Incubation of [4-32P]PtdIns4P with membranes yields only radioactive inorganic phosphate, indicating the presence of a phosphomonoesterase. 3. Membranes hydrolyze Ptd[2-3H]Ins4P to produce mainly Ptd[2-3H]Ins in the lipid phase. In this incubation [3H]inositol and inositol monophosphate appear in the water phase. 4. Membrane incubations of Ptd[2-3H]Ins4P carried out in the presence of excess non-radioactive Ins(1,4)P2 allows the trapping of small amounts of [3H]Ins(1,4)P2. These results demonstrate the presence of a phospholipase C. 5. Testing several phosphorylated analogs, it is determined that fructose 1,6-bisphosphate and alpha-glycerophosphate are potent inhibitors of the oocyte PtdIns4P phosphomonoesterase.  相似文献   

6.
D-Myo-inositol 1,4,5-trisphosphate (Ins[1,4-,5]P3) inhibits rat heart sarcolemmal Ca(2+)-ATPase activity (T. H. Kuo, Biochem. Biophys. Res. Commun. 152: 1111, 1988). We have studied the effect and mechanism of action of Ins(1,4,5)P3 and related inositol phosphates on human red cell membrane Ca(2+)-ATPase (EC 3.6.1.3) activity in vitro. At 10(-6) M, Ins(1,4,5)P3 and D-myo-inositol 4,5-bisphosphate (Ins[4,5]P2) inhibited human erythrocyte membrane Ca(2+)-ATPase activity in vitro by 42 and 31%, respectively. D-Myo-inositol 1,3,4,5-tetrakisphosphate, D-myo-inositol 1,4-bisphosphate, and D-myo-inositol 1-phosphate were not inhibitory. Enzyme inhibition by Ins(1,4,5)P3 was blocked by heparin. Exogenous purified calmodulin also stimulated red cell membrane Ca(2+)-ATPase activity; this stimulation was inhibited by Ins(1,4,5)P3. Ins(4,5)P2 and Ins(1,4,5)P3, but not Ins(1,4)P2, inhibited the binding of [125I]calmodulin to red cell membranes. Thus, specific inositol phosphates reduce plasma membrane Ca(2+)-ATPase activity and enhancement of the latter in vitro by purified calmodulin. The mechanism of these effects may in part relate to inhibition by inositol phosphates of binding of calmodulin to erythrocyte membranes.  相似文献   

7.
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   

8.
Phosphatidylinositol (PtdIns), phosphatidylinositol 4-phosphate (PtdIns4P) and phosphatidylinositol 4,5-bisphosphate [PtdIns(4,5)P2] of turkey erythrocytes were labelled by using either [32P]Pi or [3H]inositol. Although there was little basal release of inositol phosphates from membranes purified from labelled cells, in the presence of guanosine 5'-[gamma-thio]triphosphate (GTP[S]) the rate of accumulation of inositol bis-, tris- and tetrakis-phosphate (InsP2, InsP3 and InsP4) was increased 20-50-fold. The enhanced rate of accumulation of 3H-labelled inositol phosphates was linear for up to 20 min; owing to decreases in 32P specific radioactivity of phosphoinositides during incubation of membranes with unlabelled ATP, the accumulation of 32P-labelled inositol phosphates was linear for only 5 min. In the absence of ATP and a nucleotide-regenerating system, no InsP4 was formed, and the overall inositol phosphate response to GTP[S] was decreased. Analyses of phosphoinositides during incubation with ATP indicated that interconversions of PtdIns to PtdIns4P and PtdIns4P to PtdIns(4,5)P2 occurred to maintain PtdIns(4,5)P2 concentrations; GTP[S]-induced inositol phosphate formation was accompanied by a corresponding decrease in 32P- and 3H-labelled PtdIns, PtdIns4P and PtdIns(4,5)P2. In the absence of ATP, only GTP[S]-induced decreases in PtdIns(4,5)P2 occurred. Since inositol monophosphate was not formed under any condition, PtdIns is not a substrate for the phospholipase C. The production of InsP2 was decreased markedly, but not blocked, under conditions where Ins(1,4,5)P3 5-phosphomonoesterase activity in the preparation was inhibited. Thus the predominant substrate of the GTP[S]-activated phospholipase C of turkey erythrocyte membranes is PtdIns(4,5)P2. Ins(1,4,5)P3 was the major product of this reaction; only a small amount of Ins(1:2-cyclic, 4,5)P3 was released. The effects of ATP on inositol phosphate formation apparently involve the contributions of two phenomena. First, the P2-receptor agonist 2-methylthioadenosine triphosphate (2MeSATP) greatly increased inositol phosphate formation and decreased [3H]PtdIns4P and [3H]PtdIns(4,5)P2 in the presence of a low (0.1 microM) concentration of GTP[S]. ATP over the concentration range 0-100 microM produced effects in the presence of 0.1 microM-GTP[S] essentially identical with those observed with 2MeSATP, suggesting that the effects of low concentrations of ATP are also explained by a stimulation of P2-receptors. Higher concentrations of ATP also increase inositol phosphate formation, apparently by supporting the synthesis of substrate phospholipids.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

9.
We explored the stereospecificity of the fructose 2,6-bisphosphate site of rabbit muscle 6-phosphofructo-1-kinase by determination of the activation constants (Ka) of several structurally locked analogues of this potent metabolic regulator. Under the assay conditions used, the Ka of fructose 2,6-bisphosphate was 0.12 microM. The most effective synthetic analogues and their Ka's were 2,5-anhydro-D-mannitol 1,6-bisphosphate (2.9 microM), 1,4-butanediol bisphosphate (6.6 microM), hexitol 1,6-bisphosphate (40 microM), and 2,5-anhydro-D-glucitol 1,6-bisphosphate (47 microM). Ten other bisphosphate compounds were much less effective as activators of the enzyme. These findings indicate that, unlike its active site, this allosteric site of 6-phosphofructo-1-kinase does not require the furanose ring. Its basic requirement seems to be a compound with two phosphate groups approximately 9 A apart. Although the free hydroxy groups of the activator do not seem to be essential, their presence enhances appreciably the affinity of the ligand for this regulatory site.  相似文献   

10.
The accumulation of inositol polyphosphates in the cerebellum in response to agonists has not been demonstrated. Guinea pig cerebellar slices prelabeled with [3H]inositol showed the following increases in response to 1 mM serotonin: At 15 s, there was a peak in 3H label in the second messenger inositol 1,4,5-trisphosphate [Ins(1,4,5)P3], decreasing to a lower level in about 1 min. The level of 3H label in the putative second-messenger inositol 1,3,4,5-tetrakisphosphate [Ins(1,3,4,5)P4] increased rapidly up to 60 s and increased slowly thereafter. The accumulation of 3H label in various inositol phosphate isomers at 10 min, when steady state was obtained, showed the following increases due to serotonin: inositol 1,3,4-trisphosphate [Ins(1,3,4)P3], eight-fold; Ins(1,3,4,5)P4, 6.4-fold; Ins(1,4,5)P3, 75%; inositol 1,4-bisphosphate [Ins(1,4)P2], 0%; inositol 3,4-bisphosphate, 100%; inositol 1-phosphate/inositol 3-phosphate, 30%; and inositol 4-phosphate, 40%. [3H]Inositol 1,3-bisphosphate was not detected in controls, but it accounted for 7.2% of the total inositol bisphosphates formed in the serotonin-stimulated samples. The fact that serotonin did not increase the formation of Ins(1,4)P2 could be due to the fact that Ins(1,4)P2 is rapidly degraded or that Ins(1,4,5)P3 is metabolized primarily by Ins(1,4,5)P3-3'kinase to form Ins(1,3,4,5)P4. In the presence of pargyline (10 microM), [3H]Ins(1,3,4,5)P4 and [3H]Ins(1,3,4)P3 levels were increased, even at 1 microM serotonin. Ketanserin (7 microM) completely inhibited the serotonin effect, indicating stimulation of serotonin2 receptors. Quisqualic acid (100 microM) also increased the levels of [3H]Ins(1,4,5)P3, [3H]Ins(1,3,4,5)P4, and [3H]Ins(1,3,4)P3, but the profile of these increases was different.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
Many cells (including angiotensin II target cells) respond to external stimuli with accelerated hydrolysis of phosphatidylinositol 4,5-bisphosphate, generating 1,2-diacylglycerol and inositol 1,4,5-trisphosphate, a rapidly diffusible and potent Ca2+-mobilizing factor. Following its production at the plasma membrane level, inositol 1,4,5-trisphosphate is believed to interact with specific sites in the endoplasmic reticulum and triggers the release of stored Ca2+. Specific receptor sites for inositol 1,4,5-trisphosphate were recently identified in the bovine adrenal cortex (Baukal, A. J., Guillemette, G., Rubin, R., Sp?t, A., and Catt, K. J. (1985) Biochem. Biophys. Res. Commun. 133, 532-538) and have been further characterized in the adrenal cortex and other target tissues. The inositol 1,4,5-trisphosphate-binding sites are saturable and present in low concentration (104 +/- 48 fmol/mg protein) and exhibit high affinity for inositol 1,4,5-trisphosphate (Kd 1.7 +/- 0.6 nM). Their ligand specificity is illustrated by their low affinity for inositol 1,4-bisphosphate (Kd approximately 10(-7) M), inositol 1-phosphate and phytic acid (Kd approximately 10(-4) M), fructose 1,6-bisphosphate and 2,3-bisphosphoglycerate (Kd approximately 10(-3) M), with no detectable affinity for inositol 1-phosphate and myo-inositol. These binding sites are distinct from the degradative enzyme, inositol trisphosphate phosphatase, which has a much lower affinity for inositol trisphosphate (Km = 17 microM). Furthermore, submicromolar concentrations of inositol 1,4,5-trisphosphate evoked a rapid release of Ca2+ from nonmitochondrial ATP-dependent storage sites in the adrenal cortex. Specific and saturable binding sites for inositol 1,4,5-trisphosphate were also observed in the anterior pituitary (Kd = 0.87 +/- 0.31 nM, Bmax = 14.8 +/- 9.0 fmol/mg protein) and in the liver (Kd = 1.66 +/- 0.7 nM, Bmax = 147 +/- 24 fmol/mg protein). These data suggest that the binding sites described in this study are specific receptors through which inositol 1,4,5-trisphosphate mobilizes Ca2+ in target tissues for angiotensin II and other calcium-dependent hormones.  相似文献   

12.
Agonist-stimulated production of phosphatidylinositol 3,4,5-trisphosphate [PtdIns(3,4,5)P3], is considered the primary output signal of activated phosphoinositide (PI) 3-kinase. The physiological targets of this novel phospholipid and the identity of enzymes involved in its metabolism have not yet been established. We report here the identification of two enzymes which hydrolyze the 5-position phosphate of PtdIns(3,4,5)P3, forming phosphatidylinositol (3,4)-bisphosphate. One of these enzymes is the 75 kDa inositol polyphosphate 5-phosphatase (75 kDa 5-phosphatase), which has previously been demonstrated to metabolize inositol 1,4,5-trisphosphate [Ins(1,4,5)P3], inositol 1,3,4,5-tetrakisphosphate [Ins(1,3,4,5)P4] and phosphatidylinositol 4,5-bisphosphate [PtdIns(4,5)P2]. We have identified a second PtdIns(3,4,5)P3 5-phosphatase in the cytosolic fraction of platelets, which forms a complex with the p85/p110 form of PI 3-kinase. This enzyme is immunologically and chromatographically distinct from the platelet 43 kDa and 75 kDa 5-phosphatases and is unique in that it removes the 5-position phosphate from PtdIns(3,4,5)P3, but does not metabolize PtdIns(4,5)P2, Ins(1,4,5)P3 or Ins(1,3,4,5)P4. These studies demonstrate the existence of multiple PtdIns(3,4,5)P3 5-phosphatases within the cell.  相似文献   

13.
[3H]Inositol ([3H]Ins) labeling of phosphoinositides was studied in rat brain cortical membranes. [3H]Ins was incorporated into a common lipid pool through both CMP-dependent and independent mechanisms. These are as follows: (1) a reverse reaction catalyzed by phosphatidyl-inositol (PtdIns) synthase, and (2) the reaction performed by the PtdIns headgroup exchange enzyme, respectively. Membrane phosphoinositides prelabeled in either CMP-dependent or independent fashions were hydrolyzed by guanosine 5'-O-(3-thiotriphosphate) (GTP gamma S)- and carbachol-stimulated phospholipase C. Unlike CMP-dependent labeling, however, CMP-independent incorporation of [3H]Ins into lipids was inhibited by 1 mM (0.04%) sodium deoxycholate. Thus, when PtdIns labeling and phospholipase C stimulation were studied in a concerted fashion, [3H]Ins was incorporated into lipids primarily through the PtdIns synthase-catalyzed reaction because of the presence of deoxycholate required to observe carbachol-stimulation of phospholipase C. Little direct breakdown of [3H]PtdIns was detected because production of myo-[3H]inositol 1-monophosphate was minimal and myo-[3H]inositol 1,4-bisphosphate was the predominant product. Although PtdIns labeling and 3H-polyphosphoinositide formation were unaffected by GTP gamma S and carbachol and had no or little lag period, GTP gamma S- and carbachol-stimulated appearance of 3H-Ins phosphates exhibited an appreciable lag (10 min). Also, flux of label from [3H]Ins to 3H-Ins phosphates was restricted to a narrow range of free calcium concentrations (10-300 nM). These results show the concerted activities of PtdIns synthase, PtdIns 4-kinase, and phospholipase C, and constitute a simple assay for guanine nucleotide-dependent agonist stimulation of phospholipase C in a brain membrane system using [3H]Ins as labeled precursor.  相似文献   

14.
Hydrolysis of inositol phosphates by plant cell extracts.   总被引:5,自引:0,他引:5       下载免费PDF全文
S K Joseph  T Esch    W D Bonner  Jr 《The Biochemical journal》1989,264(3):851-856
A gel-filtered soluble fraction prepared from suspension-cultured Nicotiana tabacum cells hydrolysed inositol mono-, bis- and tris-phosphates. At a concentration of 7.5 microM the rates of hydrolysis followed the sequence Ins(1,4,5)P3 greater than Ins(1,4)P2 greater than Ins(4)P congruent to Ins(1)P. The major products of Ins(1,4,5)P3 hydrolysis identified by h.p.l.c. were Ins(1,4)P2 and Ins(4,5)P2. Ins(1,4)P2 was hydrolysed exclusively to Ins(4)P. The inclusion of Ca2+ in the incubation buffer markedly stimulated the hydrolysis of all the inositol phosphate substrates. Under identical conditions, Ca2+ inhibited the hydrolysis of inositol phosphates by soluble extracts prepared from rat brain. Half-maximal stimulation of Ins(1,4)P2 hydrolysis was obtained at free [Ca2+] of 0.6 and 1.2 microM when the Mg2+ concentration in the incubations was 0.3 and 1.0 mM respectively. This effect of Ca2+ was exerted solely by increasing the Vmax. of hydrolysis without affecting the Km for Ins(1,4)P2. Again, in contrast with brain, the hydrolysis of inositol bis- or mono-phosphates was insensitive to high concentrations of Li+. We conclude that plants contain specific Li+-insensitive inositol phosphate phosphatases that are regulated by low concentrations of Ca2+ in a manner which is different from that observed in mammalian tissues.  相似文献   

15.
Turner WL  Plaxton WC 《Planta》2003,217(1):113-121
Pyrophosphate-dependent phosphofructokinase (PFP; EC 2.7.1.90) and two isoforms of ATP-dependent phosphofructokinase (PFK I and PFK II; EC 2.7.1.11) from ripened banana ( Musa cavendishii L. cv. Cavendish) fruits were resolved via hydrophobic interaction fast protein liquid chromatography (FPLC), and further purified using anion-exchange and gel filtration FPLC. PFP was purified 1,158-fold to a final specific activity of 13.9 micromol fructose 1,6-bisphosphate produced (mg protein)(-1) x min(-1). Gel filtration FPLC and immunoblot analyses indicated that this PFP exists as a 490-kDa heterooctomer composed of equal amounts of 66- (alpha) and 60-kDa (beta) subunits. PFP displayed hyperbolic saturation kinetics for fructose 6-phosphate (Fru 6-P), PPi, fructose 1,6-bisphosphate, and Pi ( K(m) values = 32, 9.7, 25, and 410 microM, respectively) in the presence of saturating (5 microM) fructose 2,6-bisphosphate, which elicited a 24-fold enhancement of glycolytic PFP activity ( K(a)=8 nM). PFK I and PFK II were each purified about 350-fold to final specific activities of 5.5-6.0 micromol fructose 1,6-bisphosphate produced (mg protein)(-1) x min(-1). Analytical gel filtration yielded respective native molecular masses of 210 and 160 kDa for PFK I and PFK II. Several properties of PFK I and PFK II were consistent with their respective designation as plastid and cytosolic PFK isozymes. PFK I and PFK II exhibited: (i) pH optima of 8.0 and 7.3, respectively; (ii) hyperbolic saturation kinetics for ATP ( K(m)=34 and 21 microM, respectively); and (iii) sigmoidal saturation kinetics for Fru 6-P ( S0.5=540 and 90 microM, respectively). Allosteric effects of phospho enolpyruvate (PEP) and Pi on the activities of PFP, PFK I, and PFK II were characterized. Increasing concentrations of PEP or Pi progressively disrupted fructose 2,6-bisphosphate binding by PFP. PEP potently inhibited PFK I and to a lesser extent PFK II ( I50=2.3 and 900 microM, respectively), while Pi activated PFK I by reducing its sensitivity to PEP inhibition. Our results are consistent with: (i) the respiratory climacteric being regulated by fine (allosteric) control of pre-existing enzymes; and (ii) primary and secondary glycolytic flux control being exerted at the levels of PEP and Fru 6-P metabolism, respectively.  相似文献   

16.
Dephosphorylation of 1D-myo-inositol 1,4-bisphosphate [Ins(1,4)P2] in rat liver is catalysed by a cytosolic phosphatase that removes the 1-phosphate group. The Km for Ins(1,4)P2 is approx. 17 microM. Li+ (100 mM) causes 50% inhibition of Ins(1,4)P2 phosphatase activity when activity is measured at the very low substrate concentration of 10 nM, but on raising the substrate concentration to 100 microM there is a greater than 10-fold increase in sensitivity to Li+, suggesting that Li+ acts mainly, but not entirely, as an uncompetitive inhibitor of Ins(1,4)P2 phosphatase. In addition, rat liver cytosol shows Li+-sensitive phosphatase activity against 1D-myo-inositol 1-,3- and 4-monophosphates. The Ins(1,4)P2 1-phosphatase and inositol monophosphatase activities all share an apparent Mr of 47 x 10(3), as determined by gel-filtration chromatography. However, the Ins(1,4)P2 1-phosphatase is more sensitive to inactivation by heat, and can be separated from inositol monophosphatase activity by anion-exchange chromatography. We conclude that rat liver cytosol contains an Ins(1,4)P2 1-phosphatase that is distinct from, but in many ways similar to, inositol monophosphatase.  相似文献   

17.
We have identified a cDNA encoding a novel inositol polyphosphate 5-phosphatase. It contains two highly conserved catalytic motifs for 5-phosphatase, has a molecular mass of 51 kDa, and is ubiquitously expressed and especially abundant in skeletal muscle, heart, and kidney. We designated this 5-phosphatase as SKIP (Skeletal muscle and Kidney enriched Inositol Phosphatase). SKIP is a simple 5-phosphatase with no other motifs. Baculovirus-expressed recombinant SKIP protein exhibited 5-phosphatase activities toward inositol 1,4,5-trisphosphate, inositol 1,3,4,5-tetrakisphosphate, phosphatidylinositol (PtdIns) 4,5-bisphosphate, and PtdIns 3,4, 5-trisphosphate but has 6-fold more substrate specificity for PtdIns 4,5-bisphosphate (K(m) = 180 microM) than for inositol 1,4, 5-trisphosphate (K(m) = 1.15 mM). The ectopic expression of SKIP protein in COS-7 cells and immunostaining of neuroblastoma N1E-115 cells revealed that SKIP is expressed in cytosol and that loss of actin stress fibers occurs where the SKIP protein is concentrated. These results imply that SKIP plays a negative role in regulating the actin cytoskeleton through hydrolyzing PtdIns 4,5-bisphosphate.  相似文献   

18.
Components of the polyphosphoinositide signalling pathway have been identified in stomatal guard cells of Commelina communis L., one of the few plant systems shown unequivocally to be capable of responding to release of inositol 1,4,5-trisphosphate in the cytoplasm by increase in cytoplasmic Ca2+. 'Isolated' epidermal strips of C. communis (in which all cells other than guard cells have been killed by treatment at low pH) were radiolabelled with myo -[2n-3H]inositol or [32P]orthophosphate for 17–18 h. The phosphoinositides and inositol phosphates were extracted. Phosphoinositides were deacylated and the head groups resolved by HPLC. The water-soluble products generated by mild periodate cleavage of HPLC-purified, deacylated lipid fractions were examined. The resulting biochemical analysis led to the identification of: PtdIns, PtdIns3 P , PtdIns4 P , PtdIns(3,4) P 2 and PtdIns(4,5) P 2. Thex inositol phosphates were resolved by HPLC. Preliminary analysis of HPLC-purified putative inositol phosphate fractions resulted in the identification of each inositol phosphate class, that is, Ins P , Ins P 2, Ins P 3, Ins P 4, Ins P 5 and InsP6. Many of these inositol phosphates occurred in different isomeric forms. The presence of 3-phosphorylated phosphoinositides suggests that they may have a role in signalling in stomatal guard cells.  相似文献   

19.
Rabbit iris smooth muscle was prelabelled with myo-[3H]inositol for 90 min and the effect of carbachol on the accumulation of inositol phosphates from phosphatidylinositol 4,5-bisphosphate [PtdIns(4,5)P2], phosphatidylinositol 4-phosphate (PtdIns4P) and phosphatidylinositol (PtdIns) was monitored with anion-exchange chromatography. Carbachol stimulated the accumulation of inositol phosphates and this was blocked by atropine, a muscarinic antagonist, and it was unaffected by 2-deoxyglucose. The data presented demonstrate that, in the iris, carbachol (50 microM) stimulates the rapid breakdown of PtdIns(4,5)P2 into [3H]inositol trisphosphate (InsP3) and diacylglycerol, measured as phosphatidate, and that the accumulation of InsP3 precedes that of [3H]inositol bisphosphate (InsP2) and [3H]inositol phosphate (InsP). This conclusion is based on the following findings. Time course experiments with myo-[3H]inositol revealed that carbachol increased the accumulation of InsP3 by 12% in 15s and by 23% in 30s; in contrast, a significant increase in InsP release was not observed until about 2 min. Time-course experiments with 32P revealed a 10% loss of radioactivity from PtdIns(4,5)P2 and a corresponding 10% increase in phosphatidate labelling by carbachol in 15s; in contrast a significant increase in PtdIns labelling occurred in 5 min. Dose-response studies revealed that 5 microM-carbachol significantly increased (16%) the accumulation of InsP3 whereas a significant increase in accumulation of InsP2 and InsP was observed only at agonist concentrations greater than 10 microM. Studies on the involvement of Ca2+ in the agonist-stimulated breakdown of PtdIns(4,5)P2 in the iris revealed the following. Marked stimulation (58-78%) of inositol phosphates accumulation by carbachol in 10 min was observed in the absence of extracellular Ca2+. Like the stimulatory effect of noradrenaline, the ionophore A23187-stimulated accumulation of InsP3 was inhibited by prazosin, an alpha 1-adrenergic blocker, thus suggesting that the ionophore stimulation of PtdIns(4,5)P2 breakdown we reported previously [Akhtar & Abdel-Latif (1978) J. Pharmacol. Exp. Ther. 204, 655-688; Akhtar & Abdel-Latif (1980) Biochem. J. 192, 783-791] was secondary to the release of noradrenaline by the ionophore. The carbachol-stimulated accumulation of inositol phosphates was inhibited by EGTA (0.25 mM) and this inhibition was reversed by excess Ca2+ (1.5 mM), suggesting that EGTA treatment of the tissue chelates extracellular Ca2+ required for polyphosphoinositide phosphodiesterase activity. K+ depolarization, which causes influx of extracellular Ca2+ in smooth muscle, did not change the level of InsP3.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

20.
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   

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