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 anaerobic metabolism of marginally lethal levels of [13C]formaldehyde by Escherichia coli (K12, MU352, CRB, and CR63) was followed in vivo by 13C NMR. The products include 1,2-propanediol. Under aeration, the 1,2-propanediol is converted to hydroxyacetone. The hydroxyacetone is reconverted to 1,2-propanediol when aeration is stopped. The process can be cycled by varying the rate of aeration.  相似文献   

2.
Administration of (13)C labeled acetates ([1-(13)C], [2-(13)C] and [1,2-(13)C(2)] to Lasiodiplodia theobromae showed the tetraketide origins of both theobroxide, a potato-tuber inducing substance [1, (1S, 2R, 5S, 6R)-3-methyl-7-oxa-bicyclo[4.1.0]hept-3-en-2,5-diol]) and its carbonyldioxy derivative [2, (1S, 4R, 5S, 6R)-7,9-dioxa-3-methyl-8-oxobicyclo [4.3.0]-2-nonene-4,5-diol]. The incorporation of acetate-derived hydrogen into 1 and 2 was studied using [2-(2)H(3), 2-(13)C]acetate. Three and one deuterium atoms were incorporated at one methyl and epoxy carbons, respectively. The observed loss of deuterium atoms from the methyl group suggests a considerable amount of exchange from the methyl group of [2-(2)H(3), 2-(13)C]acetate during biosynthesis of 1 and 2. Incorporation of [1-(13)C]- and [1,2-(13)C(2)]acetates indicates the carbonyl carbon of the carbonyldioxy derivative is derived from the carboxy carbon of the precursor.  相似文献   

3.
An incorporation study of [1-(13)C] and [1,2-(13)C2] labeled sodium acetates into sorbicillinol 1 established a ring closure system between C-1 and C-6 and the positions that were oxidized and/or methylated on a hexaketide chain. Subsequent investigations, using 13C-labeled 1 prepared from [1-(13)C] labeled sodium acetate, clearly demonstrated that both bisorbicillinol 2 and sorbicillin 6 incorporated 13C-labeled 1 into their carbon skeletons. 13C-labeled bisorbicillinols 2 derived from [1-(13)C]- and [2-(13)C]-labeled sodium acetates clearly indicate that these were on the biosynthetic route from 1 to bisorbibutenolide (bislongiquinolide) 3 and bisorbicillinolide 4 via 2 as a branching point in the fungus.  相似文献   

4.
T Tamura  M Wada  N Esaki    K Soda 《Journal of bacteriology》1995,177(9):2265-2269
Streptomyces cattleya produces fluoroacetate and 4-fluorothreonine from inorganic fluoride added to the culture broth. We have shown by 19F nuclear magnetic resonance (NMR) spectrometry that fluoroacetate is accumulated first in the culture broth and that accumulation of 4-fluorothreonine is next. To show precursors of the carbon skeleton of fluoroacetate, we carried out tracer experiments with various 14C- and 13C-labeled compounds. Radioactivity of [U-14C]glucose, [U-14C]glycerol, [U-14C]serine, and [U-14C]beta-hydroxypyruvate was incorporated into fluoroacetate to an extent of 0.2 to 0.4%, whereas [3-14C]pyruvate, [2,3-14C]succinate, and [U-14C]aspartate were less efficiently incorporated (0.04 to 0.08%). The addition of [2-13C]glycerol to the mycelium suspension of Streptomyces cattleya caused exclusive enrichment of the carboxyl carbon of fluoroacetate with 13C; about 40% of carboxyl carbon of fluoroacetate was labeled with 13C. We studied the radioactivity incorporation of [3-14C]-, [U-14C]-, and [1-14C]beta-hydroxypyruvates to show that C-2 and C-3 of beta-hydroxypyruvate are exclusively converted to the carbon skeleton of fluoroacetate. These results suggest that the carbon skeleton of fluoroacetate derives from C-1 and C-2 of glycerol through beta-hydroxypyruvate, whose hydroxyl group is eventually replaced by fluoride.  相似文献   

5.
Fermentation of glycerol to 1,3-propanediol: use of cosubstrates   总被引:16,自引:0,他引:16  
Three fermentable substances, glucose, 1,2-ethanediol and 1,2-propanediol were checked as cosubstrates for the fermentation of glycerol by Clostridium butyricum and Citrobacter freundii with the aim of achieving a complete conversion of glycerol to 1,3-propanediol. Glucose was fermented by C. butyricum mainly to acetate, CO2 and reducing equivalents in the presence of glycerol and contributed markedly to the 1,3-propanediol yield. However, because of relatively slow growth on glucose, complete conversion was not achieved. If the two glycols were used as cosubstrates for glycerol fermentation, the 1,3-propanediol yield did not increase but dimished considerably, as they were converted to more reduced products, i.e. alcohols instead of acids. From 1,2-propanediol 2-propanol was formed in addition to 1-propanol. The ratio of the propanols was dependent on the culture conditions.  相似文献   

6.
Clostridium acetobutylicum is not able to grow on glycerol as the sole carbon source since it cannot reoxidize the excess of NADH generated by glycerol catabolism. Nevertheless, when the pSPD5 plasmid, carrying the NADH-consuming 1,3-propanediol pathway from C. butyricum VPI 3266, was introduced into C. acetobutylicum DG1, growth on glycerol was achieved, and 1,3-propanediol was produced. In order to compare the physiological behavior of the recombinant C. acetobutylicum DG1(pSPD5) strain with that of the natural 1,3-propanediol producer C. butyricum VPI 3266, both strains were grown in chemostat cultures with glycerol as the sole carbon source. The same "global behavior" was observed for both strains: 1,3-propanediol was the main fermentation product, and the qH2 flux was very low. However, when looking at key intracellular enzyme levels, significant differences were observed. Firstly, the pathway for glycerol oxidation was different: C. butyricum uses a glycerol dehydrogenase and a dihydroxyacetone kinase, while C. acetobutylicum uses a glycerol kinase and a glycerol-3-phosphate dehydrogenase. Secondly, the electron flow is differentially regulated: (i) in C. butyricum VPI 3266, the in vitro hydrogenase activity is 10-fold lower than that in C. acetobutylicum DG1(pSPD5), and (ii) while the ferredoxin-NAD+ reductase activity is high and the NADH-ferredoxin reductase activity is low in C. acetobutylicum DG1(pSPD5), the reverse is observed for C. butyricum VPI 3266. Thirdly, lactate dehydrogenase activity is only detected in the C. acetobutylicum DG1(pSPD5) culture, explaining why this microorganism produces lactate.  相似文献   

7.
1. Rat liver slices were employed to study the relative rates of incorporation of a mixture of [2-(3)H]- or [1,3-(3)H]-glycerol and [1-(14)C]glycerol into lipids. 2. With 0.1mm-glycerol approx. 82% of the newly synthesized lipid, calculated from (14)C incorporation, was present as neutral lipid, 13% as phosphatidylcholine and 5% as phosphatidylethanolamine. Increasing the glycerol concentration to 40mm caused a decrease in the percentage of neutral lipid to 59% and a corresponding increase in the percentage of phosphatidylcholine to 36% of the newly synthesized lipid. 3. The (d.p.m. of 2-(3)H)/(d.p.m. of 1-(14)C) ratio in glycerolipid was considerably higher than that in precursor glycerol throughout the range of experimental conditions. In contrast the incorporation of a mixture of [1,3-(3)H]glycerol and [1-(14)C]glycerol into lipid occurred with little or no change in the (3)H/(14)C ratio. 4. Respiring rat liver mitochondria were found to oxidize a mixture of sn-[2-(3)H]- and sn-[1-(14)C]-glycerol 3-phosphate with a resultant increase in the (3)H/(14)C ratio of the remaining sn-glycerol 3-phosphate. This increase is due to a (3)H isotope effect of the mitochondrial sn-glycerol 3-phosphate dehydrogenase (EC 1.1.99.5), which discriminates against sn-[2-(3)H]glycerol 3-phosphate during oxidation. 5. A method is described for the simultaneous determination of the relative contributions of the glycerol phosphate and dihydroxyacetone phosphate pathways of glycerolipid biosynthesis in rat liver slices. The method involves measurement of the (d.p.m. of 2-(3)H)/(d.p.m. of 1-(14)C) ratio in both sn-glycerol 3-phosphate and glycerolipid after incubation of rat liver slices with a mixture of [2-(3)H]glycerol and [1-(14)C]glycerol for various times. 6. By using this method it was shown that 40-50% of the glycerol incorporated into lipid by rat liver slices proceeded via the sn-glycerol 3-phosphate pathway and 50-60% was incorporated via dihydroxyacetone phosphate.  相似文献   

8.
An incorporation study of [1-13C] and [1,2-13C2] labeled sodium acetates into sorbicillinol 1 established a ring closure system between C-1 and C-6 and the positions that were oxidized and/or methylated on a hexaketide chain. Subsequent investigations, using 13C-labeled 1 prepared from [1-13C] labeled sodium acetate, clearly demonstrated that both bisorbicillinol 2 and sorbicillin 6 incorporated 13C-labeled 1 into their carbon skeletons. 13C-labeled bisorbicillinols 2 derived from [1-13C]- and [2-13C]-labeled sodium acetates clearly indicate that these were on the biosynthetic route from 1 to bisorbibutenolide (bislongiquinolide) 3 and bisorbicillinolide 4 via 2 as a branching point in the fungus.  相似文献   

9.
Glycerol has become an attractive carbon source in the biotechnology industry owing to its low price and reduced state. However, glycerol is rarely used as a carbon source in Saccharomyces cerevisiae because of its low utilization rate. In this study, we used glycerol as a main carbon source in S. cerevisiae to produce 1,2-propanediol. Metabolically engineered S. cerevisiae strains with overexpression of glycerol dissimilation pathway genes, including glycerol kinase (GUT1), glycerol 3-phosphate dehydrogenase (GUT2), glycerol dehydrogenase (gdh), and a glycerol transporter gene (GUP1), showed increased glycerol utilization and growth rate. More significant improvement of glycerol utilization and growth rate was accomplished by introducing 1,2-propanediol pathway genes, mgs (methylglyoxal synthase) and gldA (glycerol dehydrogenase) from Escherichia coli. By engineering both glycerol dissimilation and 1,2-propanediol pathways, the glycerol utilization and growth rate were improved 141% and 77%, respectively, and a 2.19 g 1,2- propanediol/l titer was achieved in 1% (v/v) glycerolcontaining YEPD medium in engineered S. cerevisiae.  相似文献   

10.
Klebsiella pneumoniae 1,3-propanediol:NAD+ oxidoreductase.   总被引:20,自引:1,他引:19       下载免费PDF全文
Fermentative utilization of glycerol, a more reduced carbohydrate than aldoses and ketoses, requires the disposal of the two extra hydrogen atoms. This is accomplished by sacrificing an equal quantity of glycerol via an auxiliary pathway initiated by glycerol dehydratase. The product, 3-hydroxypropionaldehyde, is then reduced by 1,3-propanediol NAD+:oxidoreductase (1,3-propanediol dehydrogenase; EC 1.1.1.202), resulting in the regeneration of NAD+ from NADH. The pathway for the assimilation of glycerol is initiated by an NAD-linked dehydrogenase. In Klebsiella pneumoniae the two pathways are encoded by the dha regulon which is inducible only anaerobically. In this study 1,3-propanediol:NAD+ oxidoreductase was purified from cells grown anaerobically on glycerol. The enzyme was immunochemically distinct from the NAD-linked glycerol dehydrogenase and was an octamer or hexamer of a polypeptide of 45,000 +/- 3,000 daltons. When tested as a dehydrogenase, only 1,3-propanediol served as a substrate; no activity was detected with ethanol, 1-propanol, 1,2-propanediol, glycerol, or 1,4-butanediol. The enzyme was inhibited by chelators of divalent cations. An enzyme preparation inhibited by alpha,alpha'-dipyridyl was reactivated by the addition of Fe2+ or Mn2+ after removal of the chelator by gel filtration. As for glycerol dehydrogenase, 1,3-propanediol oxidoreductase is apparently inactivated by oxidation during aerobic metabolism, under which condition the enzyme becomes superfluous.  相似文献   

11.
Glycerol kinase activity and glycerol utilization by rat granular pneumocytes were determined in order to investigate the rate-limiting step for glycerol incorporation into lung lipids. Granular pneumocytes were isolated in primary culture following trypsinization of rat lungs. Glycerol kinase activity was 8.2 nmol/h per 10(6) cells. Incorporation of [1,3-14C]glycerol into total cell lipids was 0.29 nmol/h per 10(6) cells. In the presence of saturating glycerol concentration, production of 3H2O from [2-3H]glycerol was 13 times greater than incorporation of [14C]glycerol into lipids. Glycerol phosphate dehydrogenase activity in isolated cells was approximately 10 times glycerol kinase activity. In the presence of 5.6 mM glucose, glycerol incorporation into lipids was decreased 79% and detritiation of glycerol was decreased 34%. This effect of glucose was due to a 25% increase in cell glycerol 3-phosphate content, resulting in dilution of the precursor pool and possible inhibition of glycerol phosphorylation. These results indicate that the relatively limited incorporation of glycerol into surfactant phospholipids by lung epithelial cells reflects the relatively high rate of glycerol 3-phosphate oxidation.  相似文献   

12.
Hepatocytes from fed rats were incubated for 120 min in the presence of alpha-D-[1,2-13C]glucose pentaacetate (1.7 mM), both D-[1,2-13C]glucose (1.7 mM) and acetate (8.5 mM), alpha-D-glucose penta[2-13C]acetate (1.7 mM), or D-[1,2-13C]glucose (8.3 mM). The amounts of 13C-enriched L-lactate and D-glucose and those of acetate and beta-hydroxybutyrate recovered in the incubation medium were comparable under the first two experimental conditions. The vast majority of D-glucose isotopomers consisted of alpha- and beta-D[1,2-13C]glucose. The less abundant single-labeled isotopomers of D-glucose were equally labeled on each C atom. The output of 13C-labeled L-lactate, mainly L-[2-13C]lactate and L-[3-13C]lactate, was 1 order of magnitude lower than that found in hepatocytes exposed to 8.3 mM D-[1,2-13C]glucose, in which case the total production of the single-labeled species of D-glucose was also increased and that of the C3- or C4-labeled hexose was lower than that of the other 13C-labeled isotopomers. In cells exposed to alpha-D-glucose penta[2-13C]acetate, the large majority of 13C atoms was recovered as [2-13C]acetate and, to a much lesser extent, beta-hydroxybutyrate labeled in position 2 and/or 4. Nevertheless, L-[2-13C]lactate, L-[3-13C]lactate, and single-labeled D-glucose isotopomers were also produced in amounts higher or comparable to those found in cells exposed to alpha-D-[1,2-13C]glucose pentaacetate. However, a modest preferential labelling of the C6-C5-C4 moiety of D-glucose, relative to its C1-C2-C3 moiety, and a lesser isotopic enrichment of the C3 (or C4), relative to that of C1 (or C6) and C2 (or C5), were now observed. These findings indicate that, despite extensive hydrolysis of alpha-D-glucose pentaacetate (1.7 mM) in the hepatocytes, the catabolism of its D-glucose moiety is not more efficient than that of unesterified D-glucose, tested at the same molar concentration (1.7 mM) in the presence of the same molar concentration of unesterified acetate (8.5 mM), and much lower than that found at a physiological concentration of the hexose (8.3 mM). The present results also argue against any significant back-and-forth interconversion of D-glucose 6-phosphate and triose phosphates, under conditions in which sizeable amounts of D-glucose are formed de novo from 13C-enriched Krebs cycle intermediates generated from either D-[1,2-13C]glucose or [2-13C]acetate.  相似文献   

13.
Procedures for the synthesis of fluorescent and radiolabeled analogues of phosphatidic acid are described. The fluorophore 7-nitrobenzo-2-oxa-1,3-diazole (NBD) was coupled to 6-amino-caproic acid and 12-aminododecanoic acid by reaction of NBD-chloride with the amino acids under mild alkaline conditions at room temperature. 1,2-Dioleoyl-sn-[U-14C]glycerol 3-phosphate was prepared by acylation of sn-[U-14C]glycerol 3-phosphate with oleic acid anhydride using dimethylaminopyridine as the catalyst. This compound was converted to 1-oleoyl-sn-[U-14C]glycerol 3-phosphate by hydrolysis with phospholipase A2. The lysophosphatidic acid was reacylated with NBD-aminocaproyl imidazole or NBD-aminododecanoyl imidazole to form the fluorescent, radiolabeled analogue of phosphatidic acid. Fluorescent, non-radiolabeled analogues of phosphatidic acid were prepared by phospholipase D hydrolysis of fluorescent phosphatidylcholine.  相似文献   

14.
We have evaluated the use of [1,2-13C2]propionate for the analysis of propionic acid metabolism, based on the ability to distinguish between the methylcitrate and methylmalonate pathways. Studies using propionate-adapted Escherichia coli MG1655 cells were performed. Preservation of the 13C-13C-12C carbon skeleton in labeled alanine and alanine-containing peptides involved in cell wall recycling is indicative of the direct formation of pyruvate from propionate via the methylcitrate cycle, the enzymes of which have recently been demonstrated in E. coli. Additionally, formation of 13C-labeled formate from pyruvate by the action of pyruvate-formate lyase is also consistent with the labeling of pyruvate C-1. Carboxylation of the labeled pyruvate leads to formation of [1,2-13C2]oxaloacetate and to multiply labeled glutamate and succinate isotopomers, also consistent with the flux through the methylcitrate pathway, followed by the tricarboxylic acid (TCA) cycle. Additional labeling of TCA intermediates arises due to the formation of [1-13C]acetyl coenzyme A from the labeled pyruvate, formed via pyruvate-formate lyase. Labeling patterns in trehalose and glycine are also interpreted in terms of the above pathways. The information derived from the [1, 2-13C2]propionate label is contrasted with information which can be derived from singly or triply labeled propionate and shown to be more useful for distinguishing the different propionate utilization pathways via nuclear magnetic resonance analysis.  相似文献   

15.
High resolution 13C NMR combined with chemical analysis were used to study the formation of metabolites from [1-13C]-labelled glucose by the salt-tolerant yeast Debaryomyces hansenii after transfer to media containing 8% NaCl. Time course spectroscopy of an aerobic cell suspension showed [1,3-13C]glycerol as the predominant end product. Perchloric acid extracts revealed additional less prominent incorporation of label into arabinitol, trehalose, glutamic acid, and alanine. The incorporation into trehalose and arabinitol showed a transient increase after shift to the high salinity medium. It is concluded that glycerol and arabinitol are the major organic solutes in D. hansenii, the production of glycerol being strongly induced by high salinity. Analysis of labelled extracts of D. hansenii after transfer to 8% NaCl media containing [1-13C]- or [6-13C]glucose, demonstrated that glucose is dissimilated via a combination of the Embden-Meyerhof-Parnas pathway and the pentose phosphate pathway, with the former playing a major role in glycerol formation and the latter in arabinitol production. The almost exclusive labelling of C5 of arabinitol from [6-13C]glucose indicates that the pathway to arabinitol proceeds via reduction of ribulose-5-phosphate.Abbreviations used NMR nuclear magnetic resonance - EMP Emden-Meyerhof-Parnas - PP pentose phosphate - GAP glyceraldehyde phosphate - DHAP dihydroxyacetone phosphate - ppm parts per million  相似文献   

16.
The effect of divalent cation ionophore, A23187, on the incorporation of [1-14C]palmitic acid, [1-14C]linoleic acid and [U-14C]glycerol into glycerolipids of polymorphonulcear leukocytes was examined. Ionophore A23187 stimulated the labeling of phosphatidic acid, phosphatidylglycerol, phosphatidylinositol, and diacylglycerol by both labeled fatty acids and glycerol. [1-14C]Palmitic acid and [1-14C]linoleic acid incorporation into phosphatidylcholine and triacylglycerol was reduced by the presence of the ionophore in the incubation medium, while [U-14C]glycerol labeling of these lipids was not significantly changed under identical conditions. These data reflect that the acylation of sn-glycerol 3-phosphate is activated, and the acylations of lysophosphatidyl-choline and endogenous diacylglycerol are inhibited in cells incubated with ionophore A23187. External calcium was not required for the ionophore effect on the incorporation of labeled fatty acids and glycerol. It is suggested that the ionophore alters the metabolism of the fatty acid and glycerol moieties of glycerolipids by changing the distribution of intracellular calcium of leukocytes.  相似文献   

17.
Summary The discrimination between the isotopes of hydrogen in the reaction catalyzed by yeast phosphoglucoisomerase is examined by NMR, as well as by spectrofluorometric or radioisotopic methods. The monodirectional conversion of D-glucose 6-phosphate to D-fructose 6-phosphate displays a lower maximal velocity with D-[2-2H]glucose 6-phosphate than unlabelled D-glucose 6-phosphate, with little difference in the affinity of the enzyme for these two substrates. About 72% of the deuterium located on the C2 of D-[1-13C,2-2H]glucose 6-phosphate is transferred intramolecularly to the C1 of D-[1-13C,1-2H]fructose 6-phosphate. The velocity of the monodirectional conversion of D-[U-14C]glucose 6-phosphate (or D-[2-3H]glucose 6-phosphate) to D-fructose 6-phosphate is virtually identical in H2O and D2O, respectively, but is four times lower with the tritiated than 14C-labelled ester. In the monodirectional reaction, the intramolecular transfer from the C2 of D-[2-3H]glucose 6-phosphate is higher in the presence of D2O than H2O. Whereas prolonged exposure of D-[1-13C]glucose 6-phosphate to D2O, in the presence of phosphoglucoisomerase, leads to the formation of both D-[1-13C,2-2H]glucose 6-phosphate and D-[1-13C,1-2H]fructose 6-phosphate, no sizeable incorporation of deuterium from D2O on the C1 of D-[1-13C]fructose 1,6-bisphosphate is observed when the monodirectional conversion of D-[1-13C]glucose 6-phosphate occurs in the concomitant presence of phosphoglucoisomerase and phosphofructokinase. The latter finding contrasts with the incorporation of hydrogen from 1H2O or tritium from 3H2O in the monodirectional conversion of D-[2-3H]glucose 6-phosphate and unlabelled D-glucose 6-phosphate, respectively, to their corresponding ketohexose esters.  相似文献   

18.
13C-nuclear magnetic resonance (NMR) spectroscopy was used to investigate the products of glycerol and acetate metabolism released by Leishmania braziliensis panamensis promastigotes and also to examine the interaction of each of these substrates with glucose or alanine. The NMR data were supplemented by measurements of the rates of oxygen consumption and substrate utilization, and of 14CO2 production from 14C-labeled substrate. Cells incubated with [2-13C]glycerol released acetate, succinate and D-lactate in addition to CO2. Cells incubated with acetate released only CO2. More succinate C-2/C-3 than C-1/C-4 was released from both [2-13C]glycerol and [2-13C]glucose, indicating that succinate was formed predominantly by CO2 fixation followed by reverse flux through part of the Krebs cycle. Some redistribution of the position of labeling was also seen in alanine and pyruvate, suggesting cycling through pyruvate/oxaloacetate/phosphoenolpyruvate. Cells incubated with combinations of 2 substrates consumed oxygen at the same rate as cells incubated with 1 or no substrate, even though the total substrate utilization had increased. When promastigotes were incubated with both glycerol and glucose, the rate of glucose consumption was unchanged but glycerol consumption decreased about 50%, and the rate of 14CO2 production from [1,(3)-14C]glycerol decreased about 60%. Alanine did not affect the rates of consumption of glucose or glycerol, but decreased 14CO2 production from these substrates by increasing flow of label into alanine. Although glucose decreased alanine consumption by 70%, it increased the rate of 14CO2 production from [U-14C]- and [l-14C]alanine by about 20%.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
The preparation of leucine and isoleucine labeled with 15N and of site-specific 13C-labeled isoleucines is described. This method is based on the induction of the biosynthetic pathways specific for branched chain amino acids in glutamic acid producing bacteria, and controlled provision of stable isotope labeled precursors. Corynebacterium glutamicum (ATCC 13032), a glutamic acid overproducer, was incubated in leucine production medium which consisted of a basal medium supplemented with [15N]ammonium sulfate, glucose, and sodium alpha-ketoisocaproate. production of L-[15N]leucine reached 138 mumol/ml at an isotopic efficiency of 90%. It was purified and checked by proton NMR and GC-MS. The electron impact (EI) spectrum showed 95 atom% enrichment. The cultivation of C. glutamicum in a similar medium containing alpha-ketobutyrate yielded L-[15N]isoleucine at a concentration of 120 mumol/ml. The GC-MS EI and chemical ionization (CI) spectra confirmed enrichment of 96 atom% 15N as that of the labeled precursors. The biosynthesis of L-[13C]isoleucine was carried out by induced cells which were transferred to a similar medium in which [2-13C]- or [3-13C]pyruvic acid replaced glucose. 13C NMR of the product isoleucine revealed single-site enrichment at C-3 or at C-3' respective to the precursor [13C]pyruvate; i.e., C-3 was labeled from [2-13C]pyruvate and C-3' from [3-13C]pyruvate. Mass spectrometric analysis confirmed that all molecules were labeled only in one carbon. This site-specific incorporation of [13C]pyruvate is contrasted with the labeling pattern obtained when producing cells were supplied with [2-13C]acetate, instead of pyruvate, when most label was incorporated into carbons 3 and 3' of the same isoleucine molecule.  相似文献   

20.
Alternate scan 13C and 31P NMR has been used to follow the metabolism of 13C-labeled substrates, in the presence and absence of insulin, in isolated perfused liver from fasted rats. Because both 31P and 13C NMR spectra are recorded almost simultaneously with this method, both phosphate metabolites and 13C-labeled metabolites are measured, noninvasively and repetitively, to give an immediate, broad survey of the hepatic response to a variety of stimuli. During the metabolism of [2-13C]pyruvate, [1,2-13C]ethanol, and NH4+, 13C-labeled glycogen increases synchronously with, and at the same rate as, the synthesis of 13C-labeled glucose; thus, glycogenesis was essentially a gluconeogenic process under our conditions and was unaltered by the presence of insulin. From the position of the 13C-labeled citrate peak observed in liver, the measurement of KD for the citrate-magnesium complex under our conditions, and the expression relating these quantities to the concentration of free Mg2+, the intracellular level of free Mg2+ is estimated to be 0.46 +/- 0.05 mM. Later administration of glucagon led to a rapid decrease in glycogen and citrate and a 44% increase in glycero-3-phosphocholine (GPC); increase in GPC is consistent with stimulation of liver phospholipase activity by glucagon. Simultaneous administration of two different 13C-labeled substrates, or one doubly labeled substrate, introduced multiplet structure arising from spin-spin interaction between labeled adjacent carbons into the peaks of several key metabolites. The 13C NMR intensity distributions within the several multiplets are used, within the context of a first-order model for fluxes into the Krebs cycle, to estimate relative fluxes under the conditions of the experiment.  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