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.
Two myo-inositol phosphates, myo-inositol 1,2,3-tris(dihydrogen phosphate) and myo-inositol 1,2,3,5-tetrakis(dihydrogen phosphate), have been synthesised in several steps from myo-inositol (in Chem. Abstr.: d-myo-inositol) in the form of their sodium salts. They were shown to prevent iron-gall-ink decay in cellulose items at the same level as phytic acid dodecasodium salt.  相似文献   

2.
The contact site A glycoprotein, a cell adhesion protein of aggregating Dictyostelium cells, was labeled with fatty acid, myo-inositol, phosphate and ethanolamine in vivo, indicating that the protein is anchored in the membrane by a lipid. This lipid was not susceptible to phosphatidyl inositol specific phospholipase C. When cleaved with nitrous acid or when subjected to acetolysis, the anchor released lipids which were different from those released from Trypanosoma variant cell surface glycoprotein, a protein with a known phosphatidyl inositol-glycan anchor. Resistance to weak and sensitivity to strong alkali indicated that the fatty acid in the contact site A glycolipid anchor was in an amide bond. On incubation with sphingomyelinase, a lipid with the chromatographic behavior of ceramide was released. These results suggest that the contact site A glycoprotein is anchored by a ceramide based lipid glycan.  相似文献   

3.
G A McPherson 《Life sciences》1990,47(17):1569-1577
The ability of bovine intestinal alkaline phosphatase (0.1-10 units/ml) to cleave myo-inositol bound phosphate moieties was examined. Paradoxically the digestion was optimal for a number of isomers at pH 5-7. It is possible that digestion at higher pH (9-10) does not proceed at maximal rates due to a conformation of the myo-inositol phosphate molecule which stabilizes the molecule against enzymatic attack. Alkaline phosphatase activity did not require the addition of added divalent cations. Moreover, several divalent cations, particularly zinc, were found to have a marked inhibitory effect. Further studies into this phenomenon suggested that some divalent cations can form insoluble complexes with myo-inositol phosphates, particularly those possessing a number of phosphate moieties, preventing the action of degradative enzymes. On the basis of these experiments we conclude that phosphate moieties can be removed from myo-inositol using relatively low concentrations of alkaline phosphatase as long as optimal incubation conditions are selected. This includes the use of a slightly acidic incubation media without the addition of divalent cations, particularly zinc.  相似文献   

4.
scyllo-Inositol is the all equatorial stereoisomer of myo-inositol. All possible 12 regioisomers of scyllo-inositol phosphate were synthesized for the first time via a scyllo-inositol benzoate intermediate, which was derived from a myo-inositol derivative. The stereoinversion of myo-inositol into scyllo-inositol was accomplished by Mitsunobu reaction of the vicinal cis-diol. The requisite intermediates, scyllo-inositol benzoates were obtained by benzoyl migration or random benzoylation, and phosphorylated to give scyllo-IPn.  相似文献   

5.
植酸酶作用机理的初探   总被引:3,自引:0,他引:3  
电喷雾电离-质谱联用仪分析结果表明,植酸酶水解植酸是以分步方式进行的,植酸酶能将植酸水解成植酸五磷酸酯至植酸一磷酸酯不同的中间产物。但最终产物主要为二磷酸肌醇,与一些同时形成的无机磷分子能与未水解的植酸以“-O-O”或“-O-”键形成多磷酸肌醇的更复杂的分子形式。  相似文献   

6.
Potentiometric, conductometric and 31P NMR titrations have been applied to study interactions between myo-inositol hexakisphosphate (phytic acid), (±)-myo-inositol 1,2,3,5-tetrakisphosphate and (±)-myo-inositol 1,2,3-trisphosphate with iron(III) ions. Potentiometric and conductometric titrations of myo-inositol phosphates show that addition of iron increases acidity and consumption of hydroxide titrant. By increasing the Fe(III)/InsP(6) ratio (from 0.5 to 4) 3 mol of protons are released per 2 mol of iron(III). At first, phytates coordinate iron octahedrally between P2 and P1,3. The second coordination site represents P5 and neighbouring P4,6 phosphate groups. Complexation is accompanied with the deprotonation of P1,3 and P4,6 phosphate oxygens. At higher concentration of iron(III) intermolecular P-O-Fe-O-P bonds trigger formation of a polymeric network and precipitation of the amorphous Fe(III)-InsP(6) aggregates. (31)P NMR titration data complement the above results and display the largest chemical shift changes at pD values between 5 and 10 in agreement with strong interactions between iron and myo-inositol phosphates. The differences in T(1) relaxation times of phosphorous atoms have shown that phosphate groups at positions 1, 2 and 3 are complexated with iron(III). The interactions between iron(III) ions and inositol phosphates depend significantly on the metal to ligand ratio and an attempt to coordinate more than two irons per InsP(6) molecule results in an unstable heterogeneous system.  相似文献   

7.
The effect of bradykinin on changes in the chemical levels of myo-inositol polyphosphates in renal papillary collecting tubules was investigated. Myo-inositol phosphate mass was determined by means of an enzymatic, fluorometric assay. Bradykinin induced increases in myo-inositol mono-, bis-, and trisphosphate which were both time and concentration dependent. Furthermore, the magnitude of the chemical levels of myo-inositol monophosphate formed were unlikely to be accounted for solely by the formation and degradation of myo-inositol trisphosphate. These observations are consistent with the concomitant hydrolysis of phosphatidylinositol and phosphatidylinositol bisphosphate. This study also confirms, in freshly isolated renal tubules, observations regarding bradykinin-induced phosphatidylinositol bisphosphate hydrolysis made previously in radiolabeled cultures.  相似文献   

8.
1H and 31P NMR spectra of a variety of phosphorylated myo-inositols have been analyzed using a Bruker WH-360 spectrometer. Proton and phosphorus chemical shifts and coupling constants are reported for myo-inositol 1-phosphate, myo-inositol 2-phosphate, myo-inositol 5-phosphate, myo-inositol 1,2-cyclic phosphate, myo-inositol 1,4-bisphosphate, myo-inositol 1,4,5-trisphosphate, and myo-inositol 1,3,4,5-tetrakisphosphate. These data provide the basis for the chemical identification and characterization of biologically relevant inositol phosphates.  相似文献   

9.
Myo-inositol monophosphatase (IMP) catalyzes the dephosphorylation of myo-inositol 3-phosphate in the last step of myo-inositol biosynthesis. IMP is also important in phosphate metabolism and is required for the biosynthesis of cell wall polysaccharides, phytic acid, and phosphatidylinositol. In Arabidopsis, IMP is encoded by VTC4. There are, however, two additional IMP candidate genes, IMPL1 and IMPL2, which have not yet been elucidated. In our genetic studies of Arabidopsis IMP genes, only the loss-of-function mutant impl2 showed embryonic lethality at the globular stage. All IMP genes were expressed in a similar manner both in the vegetative and reproductive organs. In developing seeds, expression of IMP genes was not coupled with the expression of the genes encoding myo-inositol phosphate synthases, which supply the substrate for IMPs in the de novo synthesis pathway. Instead, expression of IMP genes was correlated with expression of the gene for myo-inositol polyphosphate 1-phosphatase (SAL1), which is involved in the myo-inositol salvage pathway, suggesting a possible salvage pathway role in seed development. Moreover, the partial rescue of the impl2 phenotype by histidine application implies that IMPL2 is also involved in histidine biosynthesis during embryo development.  相似文献   

10.
When aleurone particles isolated from rice grains were incubated with 32P-orthophosphate or 3H-myo-inositol, both radioactivities were incorporated into an acid-stable phosphate ester. As the reaction product, myo-inositol monophosphate was recognized by ion exchange column chromatography. The phosphorylation activity was highest at the bran which corresponded to the aleurone layer. These observations suggest that the phosphorylation site of myo-inositol in the rice grain is the aleurone particles.

The phosphorylation of myo-inositol was enhanced by the existence of ATP. The optimum pH and temperature for the phosphorylation were 7.9 and 30°C, respectively.  相似文献   

11.
Phytic acid, myo-inositol-1,2,3,4,5,6-hexakisphosphate or Ins P6, is the most abundant myo-inositol phosphate in plant cells, but its biosynthesis is poorly understood. Also uncertain is the role of myo-inositol as a precursor of phytic acid biosynthesis. We identified a low-phytic acid mutant, lpa3, in maize. The Mu-insertion mutant has a phenotype of reduced phytic acid, increased myo-inositol and lacks significant amounts of myo-inositol phosphate intermediates in seeds. The gene responsible for the mutation encodes a myo-inositol kinase (MIK). Maize MIK protein contains conserved amino acid residues found in pfkB carbohydrate kinases. The maize lpa3 gene is expressed in developing embryos, where phytic acid is actively synthesized and accumulates to a large amount. Characterization of the lpa3 mutant provides direct evidence for the role of myo-inositol and MIK in phytic acid biosynthesis in developing seeds. Recombinant maize MIK phosphorylates myo-inositol to produce multiple myo-inositol monophosphates, Ins1/3P, Ins4/6P and possibly Ins5P. The characteristics of the lpa3 mutant and MIK suggest that MIK is not a salvage enzyme for myo-inositol recycling and that there are multiple phosphorylation routes to phytic acid in developing seeds. Analysis of the lpa2/lpa3 double mutant implies interactions between the phosphorylation routes.  相似文献   

12.
Microinjection of myo-inositol 1,4,5-trisphosphate into voltage-clamped Xenopus laevis oocytes or the stimulation of the phosphatidylinositol cycle elicits a complex Ca2(+)-dependent Cl- current pattern. Microinjection of myo-inositol 1,3,4,5-tetrakisphosphate causes an immediate release of Ca2+, but elicits a different Cl- current pattern than myo-inositol 1,4,5-trisphosphate. We have studied the effects of myo-inositol 1,4,5-trisphosphorothioate, which can not be converted to myo-inositol 1,3,4,5-tetrakisphosphate. Myo-inositol 1,4,5-trisphosphorothioate caused an immediate release of intracellular Ca2+, as measured by fura-2 imaging. Myo-inositol 1,4,5-trisphosphorothioate generated a Cl- current pattern similar to myo-inositol 1,3,4,5-tetrakisphosphate, not myo-inositol 1,4,5-trisphosphate.  相似文献   

13.
Detergent extraction of brain slices and mouse fibroblast 3T3 cells was performed to determine rates and relative amounts of extraction of inositol versus the glycolytic enzymes. The two detergents, Triton X-100 and Brij 58, led to similar results for extraction of myo-inositol. The extraction of enzymes from brain slices or cells varied with the detergent. In brain slices, a buffered solution containing 0.2% of the detergent Brij 58 led to the extraction of 85% of the inositol before 3% of the aldolase or before 37% of either lactate dehydrogenase or triose phosphate isomerase was extracted. In contrast, with 0.1% Triton X-100 in isotonic phosphate-buffered saline, when 70% of the inositol was extracted, 33% of the aldolase and 48% of the triose phosphate isomerase were extracted. Lesser amounts of aldolase and glyceraldehyde phosphate dehydrogenase were extracted than most of the other glycolytic enzymes under all conditions, implying that these enzymes may be interacting with non-extractable subcellular components. In 3T3 cells, both detergents were of similar effectiveness for inositol extraction. Triton X-100 caused 89% of the inositol to be released and Brij 58 caused 84% to be released. With the enzymes, Brij 58 caused between 15 and 38% extraction and Triton X-100 caused between 61 and 85% extraction of the different glycolytic enzymes. Thus Brij 58 was as effective as Triton X-100 in inositol extraction but not nearly as effective in glycolytic enzyme extraction. The results demonstrate that inositol leakage from tissues or cells is a better indicator of detergent-mediated alterations in membrane porosity than glycolytic enzyme leakage.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
A single, recessive mutation in soybean (Glycine max L. Merr.), which confers a seed phenotype of increased inorganic phosphate, decreased phytic acid, and a decrease in total raffinosaccharides, has been previously disclosed (S.A. Sebastian, P.S. Kerr, R.W. Pearlstein, W.D. Hitz [2000] Soy in Animal Nutrition, pp 56-74). The genetic lesion causing the multiple changes in seed phenotype is a single base change in the third base of the codon for what is amino acid residue 396 of the mature peptide encoding a seed-expressed myo-inositol 1-phospate synthase gene. The base change causes residue 396 to change from lysine to asparagine. That amino acid change decreases the specific activity of the seed-expressed myo-inositol 1-phosphate synthase by about 90%. Radio tracer experiments indicate that the supply of myo-inositol to the reaction, which converts UDP-galactose and myo-inositol to galactinol is a controlling factor in the conversion of total carbohydrate into the raffinosaccharides in both wild-type and mutant lines. That same decrease in myo-inositol 1-phosphate synthetic capacity leads to a decreased capacity for the synthesis of myo-inositol hexaphosphate (phytic acid) and a concomitant increase in inorganic phosphate.  相似文献   

15.
The autohydrolysis of phytic acid at 120 degrees C resulted in the formation of most of the phosphate esters of myo-inositol in varying amounts depending upon the reaction time. Eighteen of the 39 chromatographically distinct myo-inositol mono-, bis-, tris-, tetrakis-, pentakis-, and hexakisphosphates have been characterized using two different HPLC systems. These myo-inositol phosphates were partially purified by preparative anion-exchange chromatography under acidic and alkaline elution conditions. The combination of these two methods provides a two-tiered chromatographic approach to the rapid and sensitive identification of inositol phosphates in complex mixtures. Identification of the products was confirmed by 1D and 2D (1)H NMR analysis. The analytical procedure was applied to the autohydrolysis of the mixture of inositol phosphates from corn steep water.  相似文献   

16.
The direct measurement of mass of inositol trisphosphate from biologic samples is described. Separation of inositol monophosphate, bisphosphate, trisphosphate, and inositol tetrakisphosphate was achieved using anion exchange chromatography with a sodium sulfate gradient. In addition, separation of the isomers of each inositol phosphate was performed using HPLC procedures. The individual inositol phosphate fractions were subsequently dephosphorylated and desalted. The myo-inositol from each fraction was then derivatized to the hexatrimethylsilyl derivative and the myo-inositol derivatives were quantified by a novel gas chromatographic analysis using the hexatrimethylsilyl derivative of chiro-inositol as an internal concentration reference. This method is a reproducible and relatively rapid procedure for the direct quantification of inositol phosphate mass which overcomes many of the problems associated with the use of radiolabeled precursors. The method is a significant improvement over existing procedures for the quantitative determination of the mass of inositol phosphate by virtue of improved recovery, sensitivity, and technical simplicity. The applicability of this method is illustrated by the quantitative determination of inositol trisphosphate in response to norepinephrine stimulation of adult canine myocytes and cerebral cortical brain slices and by measurement of the isomers of inositol trisphosphate in isolated myocytes.  相似文献   

17.
We previously have shown an increase in muscarinic receptor density in streptozotocin (STZ)-induced diabetic and sucrosefed diuretic rat detrusor that correlates with an increase in the contractile response to muscarinic agonist (J Pharmacol Exp Ther 248: 81, 1989; Diabetes 40: 265, 1991). To investigate the signal transduction pathway involved in this altered functional response, we examined muscarinic receptor-coupled phosphatidylinositol metabolism in STZ-diabetic, sucrose-fed diuretic and age-matched control rat bladders. [3H]myo-inositol uptake was similar in all groups, but incorporation of myo-inositol into phosphatidylinositol (PI) was significantly increased in the diabetic bladder compared to the sucrose-fed and control rat bladders. Carbachol-induced increase in inositol phosphate (IPs) production was higher in the diabetic bladder than in bladders from control and sucrose-fed animals although the EC50 values were similar for all groups. Enhanced inositol phosphate production after muscarinic agonist stimulation may be due not only to the upregulation of muscarinic receptors but also to the increased incorporation of myo-inositol into PI in the STZ-induced diabetic bladder.  相似文献   

18.
The metabolism of phosphoinositides in rat sciatic nerves in vivo during electrical stimulation was studied. Nerves were prelabeled by injection of [2-3H]-myo-inositol alone for periods of 2 and 20 h or together with [32P]orthophosphate for 2 h and then electrically stimulated (100 Hz) for 5 or 20 min. Contralateral unstimulated nerve served as the control. When tritiated myo-inositol was used alone for prelabeling the nerves, approximately 6% and 14% of the label was incorporated into lipids after 2 h and 20 h, respectively. Both 5 and 20 min of electrical stimulation caused an insignificant change in the percentage of radioactivity recovered in lipids from the nerves prelabeled with either myo-inositol or with a mixture of myo-inositol and phosphate. The proportion of label associated with phosphoinositides of nerves prelabeled with myo-inositol for both 2 h and 20 h showed an increase in phosphatidyl-inositol-4-phosphate at the expense of phosphatidylinositol in stimulated nerves. Similar results were obtained with nerves prelabeled for 2 h with a mixture of [32P]orthophosphate and [2-3H]myo-inositol. No significant changes in the radioactivity associated with water-soluble inositol phosphates were found in stimulated versus control nerves.  相似文献   

19.
BACKGROUND: Phytases hydrolyze phytic acid (myo-inositol-hexakisphosphate) to less-phosphorylated myo-inositol derivatives and inorganic phosphate. Phytases are used in animal feed to reduce phosphate pollution in the environment. Recently, a thermostable, calcium-dependent Bacillus phytase was identified that represents the first example of the beta propeller fold exhibiting phosphatase activity. We sought to delineate the catalytic mechanism and property of this enzyme. RESULTS: The crystal structure of the enzyme in complex with inorganic phosphate reveals that two phosphates and four calcium ions are tightly bound at the active site. Mutation of the residues involved in the calcium chelation results in severe defects in the enzyme's activity. One phosphate ion, chelating all of the four calcium ions, is close to a water molecule bridging two of the bound calcium ions. Fluoride ion, which is expected to replace this water molecule, is an uncompetitive inhibitor of the enzyme. The enzyme is able to hydrolyze any of the six phosphate groups of phytate. CONCLUSIONS: The enzyme reaction is likely to proceed through a direct attack of the metal-bridging water molecule on the phosphorous atom of a substrate and the subsequent stabilization of the pentavalent transition state by the bound calcium ions. The enzyme has two phosphate binding sites, the "cleavage site", which is responsible for the hydrolysis of a substrate, and the "affinity site", which increases the binding affinity for substrates containing adjacent phosphate groups. The existence of the two nonequivalent phosphate binding sites explains the puzzling formation of the alternately dephosphorylated myo-inositol triphosphates from phytate and the hydrolysis of myo-inositol monophosphates.  相似文献   

20.
The inositol phosphate products formed during the cleavage of phosphatidylinositol by phosphatidylinositol-specific phospholipase C from Bacillus cereus were analyzed by 31P NMR. 31P NMR spectroscopy can distinguish between the inositol phosphate species and phosphatidylinositol. Chemical shift values (with reference to phosphoric acid) observed are 0.41, 3.62, 4.45, and 16.30 ppm for phosphatidylinositol, myo-inositol 1-monophosphate, myo-inositol 2-monophosphate, and myo-inositol 1,2-cyclic monophosphate, respectively. It is shown that under a variety of experimental conditions this phospholipase C cleaves phosphatidylinositol via an intramolecular phosphotransfer reaction producing diacylglycerol and D-myo-inositol 1,2-cyclic monophosphate. We also report the new and unexpected observation that the phosphatidylinositol-specific phospholipase C from B. cereus is able to hydrolyze the inositol cyclic phosphate to form D-myo-inositol 1-monophosphate. The enzyme, therefore, possesses phosphotransferase and cyclic phosphodiesterase activities. The second reaction requires thousandfold higher enzyme concentrations to be observed by 31P NMR. This reaction was shown to be regiospecific in that only the 1-phosphate was produced and stereospecific in that only D-myo-inositol 1,2-cyclic monophosphate was hydrolyzed. Inhibition with a monoclonal antibody specific for the B. cereus phospholipase C showed that the cyclic phosphodiesterase activity is intrinsic to the bacterial enzyme. We propose a two-step mechanism for the phosphatidyl-inositol-specific phospholipase C from B. cereus involving sequential phosphotransferase and cyclic phosphodiesterase activities. This mechanism bears a resemblance to the well-known two-step mechanism of pancreatic ribonuclease, RNase A.  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