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 pH dependence of the chemical shifts of the 31P resonances of enzyme-bound substrates 2-phosphoglycerate (PGA) and phosphoenolpyruvate (PEP) were measured to obtain further insight into the catalytic mechanism of yeast enolase. The 31P resonances of PGA and PEP bound to the enolase-Mg complex are individually observed by NMR. The Keq,internal = 1.5 favoring PEP was measured. A pH dependence of the 31P chemical shifts gives pKa values of 5.82 and 6.16 for bound PGA and PEP, respectively, indicating that both ligands bind predominantly with their phosphate groups as the dianionic species and their ionization has been altered. The phosphoryl group of PGA has been suggested as playing a role in catalysis [Nowak, T., Mildvan, A. S., and Kenyon, G. L. (1973) Biochemistry 12, 1690-1701]. The pH dependence of the kinetic parameters for Mg-enolase shows a single break in the plot of pKm, PGA vs pH at pH 6.27 with a pH independence above pH 7. This is consistent with the trianion of PGA preferably binding to the enzyme. The kcat profile gives pKA values of 5.94 and 8.35, and kcat/Km profiles give pKA values of 5.85, 6.25, and 8.39. Activation studies with Mg2+ show a pH independence for the activator constant (Ka), but a pH-dependent inhibition at higher concentrations of Mg2+. The log kcat and kcat/Ka profiles from Mg2+ activation give pKA values of about 5.9 and 8.4. These results confirm the importance of residues with pKA values of about 5.9 and 8.4 (His and Lys residues?) but do not support a function for the phosphoryl group of the substrate. The pH dependence of the Ki,Mg2+ gives pKA fits of 5. 95, 7.13, and 8.35. Data from cation inhibition suggest that the phosphate of the substrate and a His residue on enolase may bind the inhibitory Mg2+.  相似文献   

2.
Diabetic metabolic change and response to aldose reductase inhibition in the Wistar rat lens were examined with phosphorus-31 (31P) nuclear magnetic resonance (NMR) spectroscopy. To avoid artifacts in sample preparation, we used freshly excised lenses and acquired NMR data for 20 min immediately after lens extraction. The results showed a diabetes-induced time-dependent loss of ATP and phosphorylcholine (PC), an increase in alpha-glycerophosphate (alpha-GP) and inorganic phosphate and the appearance of sorbitol-3-phosphate (S-3-P) and fructose-3-phosphate (F-3-P). Oral but not topical dosing of an aldose reductase inhibitor, 5-(3-ethoxy-4-pentyloxyphenyl)-2,4- thiazolidinedione, resulted in a positive dose-response correlation characterized by a restoration of PC, S-3-P and F-3-P to the prediabetic level; however, alpha-GP and ATP were only partially normalized. The significance of the 31P change was further discussed.  相似文献   

3.
The sugar-induced inhibition of malolactic fermentation in cell suspensions of Leuconostoc oenos, recently reclassified as Oenococcus oeni (L. M. T. Dicks, F. Dellaglio, and M. D. Collins, Int. J. Syst. Bacteriol. 45:395-397, 1995) was investigated by in vivo and in vitro nuclear magnetic resonance (NMR) spectroscopy and manometric techniques. At 2 mM, glucose inhibited malolactic fermentation by 50%, and at 5 mM or higher it caused a maximum inhibitory effect of ca. 70%. Galactose, trehalose, maltose, and mannose caused inhibitory effects similar to that observed with glucose, but ribose and 2-deoxyglucose did not affect the rate of malolactic activity. The addition of fructose or citrate completely relieved the glucose-induced inhibition. Glucose was not catabolized by permeabilized cells, and inhibition of malolactic fermentation was not observed under these conditions. 31P NMR analysis of perchloric acid extracts of cells obtained during glucose-malate cometabolism showed high intracellular concentrations of glucose-6-phosphate, 6-phosphogluconate, and glycerol-3-phosphate. Glucose-6-phosphate, 6-phosphogluconate, and NAD(P)H inhibited the malolactic activity in permeabilized cells or cell extracts, whereas NADP+ had no inhibitory effect. The purified malolactic enzyme was strongly inhibited by NADH, whereas all the other above-mentioned metabolites exerted no inhibitory effect, showing that NADH was responsible for the inhibition of malolactic activity in vivo. The concentration of NADH required to inhibit the activity of the malolactic enzyme by 50% was ca. 25 microM. The data provide a coherent biochemical basis to understand the glucose-induced inhibition of malolactic fermentation in L. oenos.  相似文献   

4.
The 46 kDa enzyme 5-enolpyruvylshikimate-3-phosphate (EPSP) synthase catalyzes the condensation of shikimate-3-phosphate (S3P) and phosphoenolpyruvate to form EPSP. The reaction is inhibited by N-(phosphonomethyl)-glycine (Glp), which in the presence of S3P, binds to EPSP synthase to form a stable ternary complex. As part of a solid-state NMR characterization of this structure, 15N labels were introduced selectively into the lysine, arginine and histidine residues of EPSP synthase and distances to a 13C label in Glp and to the 31P in S3P and Glp were measured by rotational-echo double-resonance NMR. Three lysine and four arginine residues are in the proximity of the phosphate group of S3P and the carboxyl and phosphonate groups of Glp. A single histidine residue is in the vicinity of the binding site (closer to Glp than to S3P) but is more distant than the lysine and arginine residues.  相似文献   

5.
1. The 129 MHz (non-proton decoupled) and 36.4 MHz (proton decoupled) 31P NMR spectra arising from unsonicated aqueous dispersions of well defined species of phospholipid have been investigated. The phospholipids employed (and the parameters varied) include phosphatidylcholine (temperature), phosphatidylethanolamine (temperature), phosphatidic acid (temperature and p2H) and phosphatidylglycerol (temperature, p2H and Ca2+ (or Mg2+)) concentration. 2. At p2H = 7 the 31 P MNR spectra arising from saturated species of phosphatidylcholine, phosphatidylethanolamine and phosphatidylglycerol become progressively broader as the temperature is reduced below the phase transition, demonstrating reduced motion in the phosphate region of the polar headgroup. 3. In the liquid crystalline state at p2H = 7 the molecular dipolar order parameters obtained for saturated species of phosphatidylcholine, phosphatidylethanolamine and phosphatidylglycerol and very similar, and are independent of the acyl chain length for species derived from lauric and myristic acid. Thus the motion in the methylene-phosphate-methylene region is similar for these different liquid crystaline phospholipid species. 4. The 31 P NMR spectra of aqueous dispersions of 14:0/14:0 phosphatidic acid display anomalous temperature and p2H dependences. The effective chemical shift anistropy (delta v CSA EFF) at 5 degrees C varies from 71 ppm at p2H = 8.5 to 38 ppm at p2H = 2.5. Further, the motion in the phosphate region is relatively insensitive to the gel or liquid crystalline nature of the hydrocarbon chains. 5. The addition of 40 mol% Ca2+ (or Mg2+) to saturated species of phosphatidylglycerol causes an increase of approx. 20 degrees C in the hydrocarbon phase transition temperature as indicated by 31 P NMR. Equimolar concentrations of Ca2+ increase the transition temperature by approx. 70 degrees C, and no 31P NMR signal could be observed for the very condensed precipitate formed below this temperature. In the liquid crystalline state the motion in the phosphate region of the polar headgroup is not significantly affected by the presence of Ca+ or Mg2+. 6. The 31P NMR spectra obtained from 18 : 1c/18 : 1c phosphatidylethanolamine are consistent with a phase transition from a lamellar to an hexagonal (HII) phase in the region 10-15 degrees C. 7. The observed narrowing of the 31 P NMR spectra of aqueous dispersions of phospholipids as the temperature is raised toward the hydrocarbon transition temperature is discussed in terms of the "pretransition" observed in calorimetric studies.  相似文献   

6.
Five oligosaccharide alpha1-phosphates and one sulfated glycopeptide have been isolated from the hemofiltrate of one patient with end-stage renal disease. Isolation of these compounds has been achieved using reverse osmosis, ion-exchange and size-exclusion chromatography and high performance liquid chromatography. The structures were predominantly elucidated by one- and two-dimensional 1H and 31P NMR spectroscopy. The chemical structures were determined to be: 1 NeuAc alpha2-3Gal alpha1-OPO3H2; 2 NeuAc alpha2-6Galbeta1-4GlcNAc alpha1-OPO3H2; 3 NeuAc alpha2-3Galbeta1-3GalNAc alpha1-OPO3H2; 4 NeuAc alpha2-3Galbeta1-3[NeuAc alpha2-6]GalNAc alpha1-OPO3H2 (proposed structure); 5 Fuc alpha1-2Galbeta1-4[Fuc alpha1-3]GlcNAc alpha1-OPO3H2; 6 HOSO3-4Fuc alpha1-6GlcNAcbeta1-NAsn. While 2 and 3 have been previously characterized as compounds of urine and hemofiltrate, the oligosaccharide alpha1-phosphates 1, 4, and 5 could be isolated--to our knowledge--for the first time from biological material. Compound 6 is the first glycopeptide reported to contain a 4-sulfated fucose residue.  相似文献   

7.
The relationships among tissue edema, lactate accumulation, and intracellular pH in middle cerebral artery (MCA)-occluded rats were investigated with multiecho 1H magnetic resonance imaging and spatially resolved metabolic images constructed by 1H and 31P nuclear magnetic resonance (NMR) chemical shift imaging (CSI). For the effective and sensitive detection of NMR signals from the brain, outer volume suppression (OVS), reduced k-space sampling and proton irradiation were incorporated into the CSI sequences. The consecutive three measurements of calculated T2 image, lactate image, and pH image which were required for 3.75 h were repeated for four cycles of 1-16 h after MCA occlusion. Tissue edema and lactate accumulation in the infarcted region were gradually and consistently increased during the 15-h observation period. In contrast, severe acidosis was already detected on the first pH image (2-4.7 h after MCA occlusion); thereafter, the degree of acidosis became milder and showed no further progression. The dissociation between the time courses of the lactate accumulation and pH decrease was clearly demonstrated by the NMR metabolic images. Acid-base balance in cerebral infarction might be affected not only by lactate production but also by complicated interactions with tissue edema and some other factors.  相似文献   

8.
A unique method for simultaneously measuring interstitial (pHe) as well as intracellular (pHi) pH in the brains of lightly anesthetized rats is described. A 4-mm microdialysis probe was inserted acutely into the right frontal lobe in the center of the area sampled by a surface coil tuned for the collection of 31P-NMR spectra. 2-Deoxyglucose 6-phosphate (2-DG-6-P) was microdialyzed into the rat until a single NMR peak was detected in the phosphomonoester region of the 31P spectrum. pHe and pHi values were calculated from the chemical shift of 2-DG-6-P and inorganic phosphate, respectively, relative to the phosphocreatine peak. The average in vivo pHe was 7.24+/-0.01, whereas the average pHi was 7.05+/-0.01 (n = 7). The average pHe value and the average CSF bicarbonate value (23.5+/-0.1 mEq/L) were used to calculate an interstitial Pco2 of 55 mm Hg. Rats were then subjected to a 15-min period of either hypercapnia, by addition of CO2 (2.5, 5, or 10%) to the ventilator gases, or hypocapnia (PCO2 < 30 mm Hg), by increasing the ventilation rate and volume. pHe responded inversely to arterial Pco2 and was well described (r2 = 0.91) by the Henderson-Hasselbalch equation, assuming a pKa for the bicarbonate buffer system of 6.1 and a solubility coefficient for CO2 of 0.031. This confirms the view that the bicarbonate buffer system is dominant in the interstitial space. pHi responded inversely and linearly to arterial PCO2. The intracellular effect was muted as compared with pHe (slope = -0.0025, r2 = 0.60). pHe and pHi values were also monitored during the first 12 min of ischemia produced by cardiac arrest. pHe decreases more rapidly than pHi during the first 5 min of ischemia. After 12 min of ischemia, pHe and pHi values were not significantly different (6.44+/-0.02 and 6.44+/-0.03, respectively). The limitations, advantages, and future uses of the combined microdialysis/31P-NMR method for measurement of pHe and pHi are discussed.  相似文献   

9.
6-Fluoropyridoxol (6-FPOL) was evaluated as a simultaneous indicator of intracellular and extracellular pH and, hence, pH gradient in perfused rat hearts. After infusion, 19F NMR spectra rapidly showed two well-resolved peaks assigned to the intracellular and extracellular compartments, and pH was calculated on the basis of chemical shift with respect to a sodium trifluoroacetate standard. To demonstrate use of this molecule, dynamic changes in myocardial pH were assessed with a time resolution of 2 min during respiratory and metabolic alkalosis or acidosis and ischemia. For a typical heart, intracellular pH (pHi) = 7.14+/-0.01 and extracellular pH (pHe) = 7.52+/-0.02. In response to metabolic alkalosis, pHi remained relatively constant and the pH gradient increased. In contrast, respiratory challenge caused a significant increase in pHi. Independent measurements using pH electrodes and 31P NMR confirmed validity of the 19F NMR results.  相似文献   

10.
BACKGROUND & AIMS: Proton pump inhibitors administered twice daily do not provide complete nocturnal acid suppression. Acid breakthrough, or decrease in intragastric pH to <4 for an hour or longer, occurs in three quarters of normal subjects and patients at night. We compared the effect of a third dose of omeprazole at bedtime with that of a dose of ranitidine at bedtime on residual nocturnal acid secretion in patients receiving omeprazole twice daily. METHODS: Twelve volunteers underwent overnight intragastric pH monitoring after 7 days of treatment with omeprazole, 20 mg twice daily, followed by different treatment supplements at bedtime: placebo; additional omeprazole, 20 mg; ranitidine, 150 mg; and ranitidine, 300 mg. RESULTS: Additional omeprazole at bedtime reduced the percentage of time with intragastric pH of <4 from 48% to 31% (P < 0.005) compared with omeprazole twice daily with placebo at bedtime. Ranitidine at bedtime reduced this parameter more, 5% with 150 mg and 6% with 300 mg (P <0.01 vs. omeprazole twice daily plus bedtime). Results for percentage of time with intragastric pH <3 were similar. Eleven subjects had acid breakthrough with placebo at bedtime; 7 with omeprazole at bedtime (P = NS); 4 with ranitidine, 150 mg at bedtime; and 3 with ranitidine, 300 mg at bedtime (P < 0. 05, ranitidine vs. placebo). CONCLUSIONS: Bedtime ranitidine is more effective than bedtime omeprazole on residual nocturnal acid secretion in patients receiving omeprazole twice daily. This finding suggests that fasting breakthrough nocturnal acid secretion in patients receiving omeprazole twice daily is most likely histamine related.  相似文献   

11.
EPSP (5-enolpyruvylshikimate-3-phosphate) synthase catalyzes condensation of shikimate 3-phosphate (S3P) and phosphoenolpyruvate (PEP) to form EPSP, a precursor to the aromatic amino acids. S3P and [2-13C]POP were bound to mutant or wild type E. coli forms of the enzyme prior to lyophilization. CPMAS-echo and rotational-echo double-resonance (REDOR) NMR experiments, employing a slow catalytic EPSP synthase mutant and a long prelyophilization incubation interval, allowed our observation of the gradual formation of a strong 31P-13C coupling consistent with the well characterized tetrahedral intermediate. However, after shorter low temperature incubation intervals of substrates with mutant or wild-type enzymes, carbon CPMAS-echo NMR spectra showed the 13C label at 155 ppm, consistent with sp2 geometry of this carbon. REDOR revealed that the phosphorus of PEP was cleaved. However, phosphorus at a distance of 7.5 A was observed, due to the phosphate of a nearby bound S3P. Heating the sample allowed the reaction to progress, as shown by the diminution of the 155 ppm peak and growth of a peak at 108 ppm. The sp3 geometry implied by the 108 ppm peak strongly suggested formation of a S3P-PEP condensation product. REDOR indicated that phosphorus was still distant, but now only 6.1 (wild type) or 5.9 A (mutant) distant. We think that the early intermediates with peaks at 155 and 108 ppm are covalently bound to the enzyme. We also think that the tetrahedral intermediate that we observed was formed after product was generated.  相似文献   

12.
The model peptides glycylglycyltyrosylalanine (Gly-Gly-Tyr-Ala), glycylglycylthreonylalanine (Gly-Gly-Thr-Ala) and glycylglycylserylalanine (Gly-Gly-Ser-Ala) were phosphorylated at the hydroxyl groups of their tyrosyl, threonyl and seryl residues, respectively, and characterized by 31P and 1H NMR spectroscopy. The pKa-value of the phosphoryl group in the tyrosine-containing peptide determined from the pH dependence of chemical shifts is 5.9, the 31P chemical shifts at low pH (4.0) and high pH (8.0) are -3.8 and 0.2 ppm, respectively. Phosphorylation also leads to significant shifts of the 1H NMR resonances of the tyrosine residue; the amide resonance is shifted -0.02 ppm, the H alpha resonance 0.06 ppm, the H beta resonances 0.10 and -0.04 ppm, the H delta resonances 0.02 ppm and the H epsilon resonances 0.26 ppm. The pKa-value of the phosphoryl group in the threonine peptide determined from the pH dependence of chemical shifts is 6.1; the 31P chemical shifts at low pH (4.0) and high pH (8.0) are -0.1 and 4.8 ppm, respectively. The corresponding values for the serine peptide are 6.1 (pKa), 0.6 ppm and 4.9 ppm. Phosphorylation also leads to significant shifts of the 1H NMR resonances of the threonine and serine residues. In the threonine residue the amide resonance is shifted 0.25 ppm, the H alpha-resonance -0.43 ppm, the H beta-resonance 0.03 ppm and the H gamma-resonance 0.09 ppm. In the serine residue the amide resonance is shifted 0.21 ppm, the H alpha-resonance -0.17 ppm, and the H beta-resonances 0.17 ppm.  相似文献   

13.
BACKGROUND: Fructose-1,6-bisphosphate (FBP) sometimes provides substantial cerebral protection during hypoxia or ischemia. 31P/1H nuclear magnetic resonance spectroscopy of cerebrocortical slices was used to study the effects of FBP on hypoxia-induced metabolic changes. In addition, 13C-labeled glucose was administered and 13C nuclear magnetic resonance spectroscopy was used to search for FBP-induced modulations in glycolysis and the pentose-phosphate pathway. METHODS: In each experiment, 80 slices (350 microm) obtained from ten 7-day-old Sprague-Dawley rat litter mates were placed together in a 20-mm nuclear magnetic resonance tube, perfused, and subjected to 30 min of hypoxia (PO2 < 3 mmHg). Nine experiments were performed, with n = 3 in each of three groups: (1) no treatment with FBP; (2) 60 min of prehypoxia treatment with FBP (2 mM); and (3) 60 min of posthypoxia treatment with FBP (2 mM). 31P/1H Interleaved nuclear magnetic resonance spectra at 4.7 T provided average adenosine triphosphate, intracellular pH, and lactate. Cresyl violet stains of random slices taken at predetermined time points were studied histologically. Some experiments had [2-13C]glucose in the perfusate. Slices from these studies were frozen for perchloric acid extraction of intracellular metabolites and studied with high-resolution 13C nuclear magnetic resonance spectroscopy at 11.75 T. RESULTS: With no pretreatment with FBP, hypoxia caused an approximately 50% loss of adenosine triphosphate, an approximately 700% increase in lactate, and a decrease in intracellular pH to approximately 6.4. Pretreatment with FBP resulted in no detectable loss of adenosine triphosphate, no increase in lactate, and minimal morphologic changes but did not alter decreases in intracellular pH. 13C Nuclear magnetic resonance spectra of extracted metabolites showed that pretreatment caused accumulation of [1-13C]fructose-6-phosphate, an early pentose-phosphate pathway metabolite. Posthypoxic treatment with FBP had no effects compared with no treatment. CONCLUSIONS: During severe hypoxia, pretreatment with FBP completely preserves adenosine triphosphate and almost completely preserves cell morphology but does not alter hypoxia-induced decreases in intracellular pH. Pretreatment also substantially augments the flux of glucose into the pentose-phosphate pathway.  相似文献   

14.
Expression of insulin-like growth factor-I (IGF-I) receptors and insulin-like growth factor-II/mannose-6-phosphate (IGF-II/Man6P) receptors in cultured bovine alveolar macrophages (BAM) was demonstrated by competitive binding studies and crosslinking experiments. Western blotting of protein extracts from cultured BAM using an anti bovine IGF-II/Man6P receptor antiserum (#66416) confirmed the presence of IGF-II/Man6P receptors on BAM. The effects of IGFs and Man6P on generation of inositol phosphates was measured by HPLC analysis of perchloric acid extracts from myo-[3H]inositol-labelled cultured BAM. IGF-I at nanomolar concentrations and Man6P (10[-8]-10[-3] M) stimulated the accumulation of both Ins(1,4,5)P3 and Ins(1,3,4,5)P4 after 30 sec. IGF-II (up to 2.3 x 10[-8] M) had no significant effect on inositol phosphate accumulation under the same conditions. Both IGFs and Man6P induced a rise in [Ca2+]i in cultured BAM. In addition, using the fluorescent dye SNARF-1/AM we could demonstrate rapid but small IGF-II (10[-9] M) triggered acidification (0.07 pH units) of cultured BAM. Taken together, our results indicate not only the presence of both IGF-I and IGF-II/Man6P receptors on BAM, but also provide evidence of the linkage of the IGF-I receptor to the inositol phosphate system.  相似文献   

15.
This randomized, single-blind cross-over study compared the effectiveness of a conventional alginate reflux barrier formulation (20 mL single dose of Liquid Gaviscon; sodium alginate, sodium bicarbonate, calcium carbonate) with a 20 mL single dose of an alginate-cimetidine combination formulation (Algitec Suspension; sodium alginate, cimetidine) in the suppression of food and acid reflux into the oesophagus after a test meal in 12 healthy volunteers. Subjects were fasted overnight before the study. A pH electrode and gamma detector were accurately positioned 5 cm above the cardia. The volunteers received a 99mTc-labelled meal designed to provoke reflux and then either remained untreated, or 30 min later were given either Algitec Suspension or Liquid Gaviscon. Reflux of both food and acid into the oesophagus was measured for 3 h. There was a seven day wash-out period between each treatment. Food reflux in the control group was 22,878 +/- 14,385 counts x 10(3) and this was significantly suppressed by both Liquid Gaviscon (174 +/- 128 (s.e.) counts x 10(3); P = 0.003); however, although the reduction of food reflux to 3812 +/- 2322 counts x 10(3) observed after Algitec treatment was considerable, this did not reach statistical significance (P > 0.05) due to the large intersubject variation. Liquid Gaviscon was significantly better at reducing food reflux than Algitec (P = 0.001). Gaviscon also significantly reduced acid reflux when compared with the control group (1.08 +/- 0.73 vs 5.87 +/- 3.27% recording time oesophageal pH < 4, respectively) (P = 0.03). The slight reduction in acid reflux after Algitec treatment (3.25 +/- 1.82% recording time oesophageal pH < 4) also did not reach statistical significance. The difference between Algitec and Gaviscon treatment was also not significant.  相似文献   

16.
The Pi peak in a 31P NMR spectrum of the brain can be deconvoluted into six separate Lorentzian peaks with the same linewidth as that of the phosphocreatine peak in the spectrum. In an earlier communication we showed that the six Pi peaks in normal brain represent two extracellular and four intracellular compartments. In that report we have identified the first of the extracellular peaks by marking plasma with infused Pi, thereby substantially increasing the amplitude of the single peak at pH 7.35. 2-Deoxyglucose-6-phosphate (2-DG-6-P) was placed in the brain interstitial space by microdialysis. The resulting 2-DG-6-P peak was deconvoluted into three separate peaks. The chemical shift of the principle 2-DG-6-P peak gave a calculated pH of 7.24 +/- 0.02 for interstitial fluid pH, a value that agreed well with the pH of the second extracellular Pi peak at pH 7.25 +/- 0.01. We identified the intracellular compartments by selectively stressing cellular energy metabolism in three of the four intracellular spaces. A seizure-producing chemical, flurothyl, was used to activate the neuron, thereby causing a demand for energy that could not be completely met by oxidative phosphorylation alone. The resulting loss of high-energy phosphate reserves caused a significant increase in intracellular Pi only in those cells associated with the Pi peak at pH 6.95 +/- 0.01. This suggests that this compartment represents the neuron. Ammonia is detoxified in the astrocyte (glutamine synthetase) by incorporating it into glutamine, a process that requires large amounts of glucose and ATP. The intraarterial infusion of ammonium acetate into the brain stressed astrocyte energy metabolism resulting in an increase in the Pi of the cells at pH of 7.05 +/- 0.01 and 7.15 +/- 0.02. This finding, coupled with our observation that these same cells take up infused Pi probably via the astrocyte end-foot processes, lead us to conclude that these two compartments represent two different types of astrocytes, probably protoplasmic and fibrous, respectively. As a result of this study, we now believe the brain contains four extracellular and four intracellular compartments.  相似文献   

17.
The substrate binding site of polyenoic fatty acid isomerase (PFI) has been investigated using a series of alternate substrates and by examination of the pH dependence on the kinetic parameters of PFI with selected substrates. The pH dependence profile of PFI with EPA [(5Z,8Z,11Z,14Z,17Z)-eicosapentaenoic acid] shows the enzyme to be catalytically active over a wide pH range, with activity being optimal below pH 6.0. Analysis of the kinetic parameters of DHA [(4Z,7Z,10Z,13Z,16Z,19Z)-docosahexen oic acid]; adrenic acid [(7Z,10Z,13Z,16Z)-docosatetraenoic acid]; EPA; arachidonic acid [(5Z,8Z,11Z,14Z)-eicosatetraenoic acid]; anandamide (arachidonyl-N-ethanolamide); and eicosatrienoic acid [(5Z,8Z,11Z)-eicosatrienoic acid] demonstrates that substrates possessing omega-3 olefins (DHA and EPA) have the lowest K(m) values (1.9 and 9.6 microM, respectively). EPA and arachidonic acid showed the highest V(max) values (6.0 and 2.8 micromol min(-1) mg(-1), respectively). The twenty carbon omega-9 fatty acid eicosatrienoic acid showed a relatively large K(m) and had a V(max) approximately 20-fold less than EPA. Anandamide, a substrate analog lacking an ionizable carboxylate, showed a K(m) similar to the other omega-6 fatty acids (arachidonic acid and adrenic acid); however, the V(max) was approximately 5-fold lower than arachidonic acid and 8-fold lower than EPA. Moreover, anandamide demonstrated no pH dependency on its kinetic parameters over a range where EPA showed a 27-fold decrease in V/K(m). NMR spectroscopy was used to determine the structure of the product from reaction of PFI with DHA. These data showed the compound to be (4Z,7Z,9E,11E,16Z,19Z)-docosahexenoi c acid. Reaction of PFI with dihomo-gamma-linolenic acid resulted in the development of two products, one with the characteristic chromophore of a conjugated triene, the other with a chromophore characteristic of a conjugated diene. Analysis of the products from these reactions of PFI, in conjunction with the kinetic parameters from the alternate substrates, provides compelling evidence that the enzyme preferentially orients the substrate in the catalytic site with respect to the methyl terminus.  相似文献   

18.
The kinetics of PI-PLCgamma1 toward a water-soluble substrate (inositol 1,2-cyclic phosphate, cIP) and phosphatidylinositol (PI) in detergent mixed micelles were monitored by 31P NMR spectroscopy. That cIP is also a substrate (Km = approximately 15 mM) implies a two-step mechanism (intramolecular phosphotransferase reaction to form cIP followed by cyclic phosphodiesterase activity to form inositol-1-phosphate (I-1-P)). PI is cleaved by PI-PLCgamma1 to form cIP and I-1-P with the enzyme specific activity and ratio of products (cIP/I-1-P) regulated by assay temperature, pH, Ca2+, and other amphiphilic additives. Cleavage of both cIP and PI by the enzyme is optimal at pH 5. The effect of Ca2+ on PI-PLCgamma1 activity is unique compared with other isozymes enzymes: Ca2+ is necessary for the activity and low Ca2+ activates the enzyme; however, high Ca2+ inhibits PI-PLCgamma1 hydrolysis of phosphoinositides (but not cIP) with the extent of inhibition dependent on pH, substrate identity (cIP or PI), substrate presentation (e.g. detergent matrix), and substrate surface concentration. This inhibition of PI-PLCgamma1 by high Ca2+ is proposed to derive from the divalent metal ion-inducing clustering of the PI and reducing its accessibility to the enzyme. Amphiphilic additives such as phosphatidic acid, fatty acid, and sodium dodecylsulfate enhance PI cleavage in micelles at pH 7.5 but not at pH 5.0; they have no effect on cIP hydrolysis at either pH value. These different kinetic patterns are used to propose a model for regulation of the enzyme. A key hypothesis is that there is a pH-dependent conformational change in the enzyme that controls accessibility of the active site to both water-soluble cIP and interfacially organized PI. The low activity enzyme at pH 7.5 can be activated by PA (or phosphorylation by tyrosine kinase). However, this activation requires lipophilic substrate (PI) present because cIP hydrolysis is not enhanced in the presence of PA.  相似文献   

19.
The driving force behind this new chromatographic technique was to develop a method for purifying preparative quantities of phosphohistidines in a single step that provided good resolution wit eluants that could be easily removed. The current method can provide milligram quantities of each phosphohistidine; however, the later 1H NMR analysis of the dried, individually purified phosphohistidines showed that histidine was present along with each of the individual phosphohistidines. The stability of each phosphohistidine during storage does not appear to be a problem because the amounts of histidine contamination of freshly freeze-dried samples were compared with those of samples stored at -80 degrees C under nitrogen for 2 weeks and were found to be similar (data not shown). Possibly, the lyophilization and preparation of solutions for NMR analysis resulted in a certain amount of hydrolysis of phosphohistidine, particularly with the less stable 1- and 1,3-forms (5, 6). It was noted that when the lyophilized samples were initially dissolved in D2O for NMR analysis, the pH was between 6 and 7; this may have resulted in some hydrolysis of the phosphohistidines. This hydrolysis can be reduced by the addition of 50 mM potassium hydroxide to the pooled fractions collected from the chromatography, so that the alkalinity of the samples is maintained throughout the subsequent processes. The data obtained for the assignment of individual phosphohistidines by 1H and 31P NMR analysis seem consistent with those obtained by other independent studies (6, 10). The NMR analysis was a powerful tool in assigning the identity of each purified phosphohistidine, although caution should be used when considering free phosphohistidines as references for NMR detection of phosphohistidines in native proteins. Lecroisey et al. (10) showed that there were differences between the chemical shifts observed for free phosphohistidine compared to those for phosphohistidine in dipeptides. However, for the purposes of phosphoamino acid analysis, these purified phosphohistidines are used by this group as references in the detection of phosphohistidine in proteins.  相似文献   

20.
A combined in vitro/in vivo study was performed to evaluate the possible application of phosphorus (31P) NMR spectroscopy for therapy monitoring and to investigate glucosylifosfamide mustard (Glc-IPM) transport and biodistribution by radiotracer techniques. Dynamic in vivo 31P NMR measurements were performed in rats with prostate adenocarcinoma after i.v. injection of 1 mmol/kg body weight (bw) of ifosfamide (IFO) (n = 4) and 1 mmol/kg bw (n = 4) or 2.15 mmol/kg bw (n = 9) of Glc-IPM. In a biodistribution study with 14C-labeled Glc-IPM and a final dose of 0.8 mmol Glc-IPM/kg bw, the animals were killed 5, 30, 60, and 120 min after drug administration, an ethanol extraction was performed from several tissues, and the dose per g tissue was calculated. The same tumor cell line was used in saturation and competition experiments to further elucidate the transport mechanism. The 31P NMR signals of IFO and Glc-IPM showed no overlap with the endogenous phosphorus peaks. A rapid washout with a half-life between 25.9 +/- 5.6 min for the lower dose and 34.3 +/- 4.2 min for the higher dose of Glc-IPM was observed in the tumor. No statistically significant change of the pH value was observed during the examination period. The beta-nucleoside 5'-triphosphate (NTP)/inorganic phosphate (Pi) signal intensity ratio showed a tendency to decrease but without statistical significance. A rapid elimination was demonstrated by both the noninvasive NMR technique and the biodistribution study. No saturation was found in vitro for the Glc-IPM uptake, even at the concentration of 5 mM. Furthermore, the Glc-IPM uptake was not inhibited by the presence of 2-deoxyglucose and vice versa. The data show that the pharmacokinetics of Glc-IPM in the tumor can be followed in vivo by 31P NMR. The results presented are evidence for diffusion as the transport mechanism for Glc-IPM in this tumor model. However, the better visualization of Glc-IPM as compared to ifosfamide may be due to metabolic trapping of a negatively charged metabolite after deglycosylation.  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