首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Threonine kinetics, threonine oxidative pathway, and the relationship between threonine and whole body protein turnover were quantified in 10 healthy term infants during the first 48 h after birth. The kinetic data were obtained 6 h after the last feed (fasting) and in response to formula feeding, using [U-(13)C(4),(15)N]threonine, [(2)H(5)]phenylalanine, and [(15)N]glycine tracers. The rate of carbon dioxide production (Vco(2)) and (13)C enrichment of the expired CO(2) were measured to quantify the rate of oxidation of threonine. The rate of appearance (R(a)) of threonine (136 +/- 37 micromol.kg(-1).h(-1)) was higher in newborn infants than that reported in adults. Formula feeding resulted in a significant decrease in threonine R(a) (P < 0.05). A significant positive correlation was seen between phenylalanine R(a) and threonine R(a), both during fasting and after formula feeding (r(2) = 0.65). In contrast to a 1:1 ratio of threonine and phenylalanine in mixed muscle protein, threonine R(a) relative to phenylalanine R(a) was 2.2 +/- 0.4. The fractional rate of threonine flux oxidized was 20% during fasting and 26% (P < 0.05) in response to nutrient administration. There was a significant correlation between plasma threonine concentration and threonine oxidation (r(2) = 0.75). No measurable incorporation of threonine in plasma glycine was seen. These data suggest that threonine is exclusively degraded by the glycine-independent serine/threonine dehydratase pathway. A higher flux of threonine relative to phenylalanine indicates higher turnover of threonine enriched proteins.  相似文献   

2.
We developed a method to measure hemoglobin synthesis rate (SynHb) in humans, assuming that free glycine in the red blood cell (RBC) represents free glycine in bone marrow for hemoglobin synthesis. The present rat study examines this assumption of the method and quantifies SynHb in rats. Sprague-Dawley rats (n = 9) were studied, [2-(13)C]glycine was intravenously infused over 24 h (2.5 mg kg(-1) h(-1)), blood was drawn for glycine and heme isolation, and bone marrow was harvested for glycine isolation. Isotopic enrichments of glycine and heme were measured, fractional hemoglobin synthesis rate (fSynHb% day(-1)) was calculated, and from this a value for SynHb (mg g(-1) day(-1)) was derived. Mean body weight was 446 +/- 10 g (mean +/- SE) and hemoglobin concentration was 14 +/- 0.5 g dl(-1). At 24 h, the mean isotopic enrichment, atom percentage excess (APE), of the RBC free glycine (1.56 +/- 0.18 APE) was similar to the bone marrow (1.68 +/- 0.15 APE). The rate of incorporation of (13)C into heme increased over time from 0.0004 APE/h between 6 and 12 h, to 0.0014 APE/h between 12 and 18 h, and 0.0024 APE/h between 18 and 24 h. Consequently, fSynHb (1.19 +/- 0.32, 2.92 +/- 0.66, and 4.22 +/- 0.56% day(-1), respectively) and SynHb (0.11 +/- 0.03, 0.28 +/- 0.05, and 0.42 +/- 0.05 mg g(-1) day(-1), respectively) showed similar patterns over the 24-h study period. We conclude that (1) enrichment of free glycine in the circulating RBC approximates enrichment of bone marrow free glycine for heme formation and (2) this pattern of hemoglobin synthesis rate is reflecting the characteristic release and gradual maturation of reticulocytes in the circulation.  相似文献   

3.
Iida K  Kajiwara M 《The FEBS journal》2007,274(19):5090-5095
The metabolic pathways leading from l-[2-13C]aspartic acid, [2-13C]glycine and l-[methyl-13C]methionine to vitamin B12 were investigated, focusing on the biosynthetic pathways leading to the aminopropanol moiety of vitamin B12 and on the role of the Shemin pathway leading to delta-aminolevulinic acid (a biosynthetic intermediate of tetrapyrrole), by means of feeding experiments with Propionibacterium shermanii in combination with 13C-NMR spectroscopy. The 13C-methylene carbons of l-[2-(13)C]aspartic acid, which is transformed to [2-13C]glycine via l-[2-13C]threonine, and [2-13C]glycine added to the culture medium served mainly to enrich the seven methyl carbons of the corrin ring through C-methylation by S-adenosyl-l-[methyl-13C]methionine derived from catabolically generated l-[methyl-13C]methionine in the presence of tetrahydrofolic acid. The results indicate that the catabolism of these amino acids predominates over pathways leading to (2R)-1-amino-2-propanol or delta-aminolevulinic acid in P. shermanii. Feeding of l-[methyl-13C]methionine efficiently enriched all seven methyl carbons. In the cases of [2-13C]glycine and l-[methyl-13C]methionine, the 13C-enrichment ratio of the methyl carbon at C-25 (the site of the first C-methylation) was less than those of the other six methyl carbons, probably due to the influence of endogenous d-glucose in P. shermanii. The almost identical 13C-enrichment ratios of the other six methyl carbons indicated that these C-methylations during vitamin B12 biosynthesis were completed before the amino acids were completely consumed. However, in the case of l-[2-13C]aspartic acid, the 13C-enrichment ratios of five methyl carbons were low and similar, whereas the last two sites of C-methylation (C-53 and C-35) were not labeled, presumably because of complete consumption of the smaller amount of added label. The ratios of 13C-incorporation into the seven methyl carbons are influenced by the conditions of amino acid feeding experiments in a manner that is dependent upon the order of C-methylation in the corrin ring of vitamin B12.  相似文献   

4.
Uptake and Release of Glycine in the Guinea Pig Cochlear Nucleus   总被引:4,自引:2,他引:2  
This study attempts to determine if the cochlear nucleus (CN) contains glycinergic synaptic endings. The uptake and release of exogenous radiolabeled glycine were measured in vitro in the three major subdivisions of the guinea pig CN: anteroventral, posteroventral, and dorsal. A kinetic analysis of [3H]glycine uptake revealed the presence in each CN subdivision of a high- and a low-affinity uptake mechanism. The high-affinity mechanism had a Km of 25.2-30.5 microM and a Vmax of 3.8-4.8 nmol/10 mg of cell water/5 min, whereas the low-affinity mechanism had a Km of 633-718 microM and a Vmax of 26.6-37.1 nmol/10 mg of cell water/5 min. At steady state, the high-affinity mechanism accumulated 10 microM [3H]glycine from the medium, achieving tissue concentrations that were 13-24 times that in the medium. The high-affinity uptake was dependent on the temperature and on the concentrations of NaCl and glucose in the incubation medium. It exhibited a high degree of substrate specificity, as determined by the effects of structural analogues of glycine on the uptake of [3H]glycine. Each CN subdivision also contained two mechanisms mediating [14C]glycine release. One was activated by depolarizing electrical stimuli, produced a rapid transient release of [14C]glycine, and was dependent on the presence of extracellular Ca2+. The other was continuous, producing a slow spontaneous efflux of [14C]glycine. Released glycine could be removed primarily by uptake, because during release measurements, the amount of [14C]glycine detected in the medium decreased when glycine uptake activity was optimized. The electrically evoked, Ca2+-dependent release and the high-affinity uptake of glycine may mediate the synaptic release and inactivation of glycine, respectively. These findings, therefore, support the presence of glycinergic synaptic endings in each CN subdivision.  相似文献   

5.
The catabolism of glycine in the isolated perfused rat liver was investigated by measuring the production of 14CO2 from [1-14C]- and [2-14C]glycine. Production of 14CO2 from [1-14C]glycine was maximal as the perfusate glycine concentration approached 10 mM and exhibited a maximal activity of 125 nmol of 14CO2 X g-1 X min-1 and an apparent Km of approximately 2 mM. Production of 14CO2 from [2-14C]glycine was much lower, approaching a maximal activity of approximately 40 nmol of 14CO2 X g-1 X min-1 at a perfusate glycine concentration of 10 mM, with an apparent Km of approximately 2.5 mM. Washout kinetic experiments with [1-14C]glycine exhibited a single half-time of 14CO2 disappearance, indicating one metabolic pool from which the observed 14CO2 production is derived. These results indicate that the glycine cleavage system is the predominant catabolic fate of glycine in the perfused rat liver and that production of 14CO2 from [1-14C]glycine is an effective monitor of metabolic flux through this system. Metabolic flux through the glycine cleavage system in the perfused rat liver was inhibited by processes which lead to reduction of the mitochondrial NAD(H) redox couple. Infusion of beta-hydroxybutyrate or octanoate inhibited 14CO2 production from [1-14C]glycine by 33 and 50%, respectively. Alternatively, infusion of acetoacetate stimulated glycine decarboxylation slightly and completely reversed the inhibition of 14CO2 production by octanoate. Metabolic conditions which are known to cause a large consumption of mitochondrial NADPH (e.g. ureogenesis from ammonia) stimulated glycine decarboxylation by the perfused rat liver. Infusion of pyruvate and ammonium chloride stimulated production of 14CO2 from [1-14C]glycine more than 2-fold. Lactate plus ammonium chloride was equally as effective in stimulating glycine decarboxylation by the perfused rat liver, while alanine plus ammonium chloride was ineffective in stimulating 14CO2 production.  相似文献   

6.
Addition of selected amino acids could be a means to improve production of recombinant proteins in industrial processes. We found that glycine increased the maximum specific growth rate of Escherichia coli from 0.67 to 0.78 h(-1), and the cell yield from 0.57 to 0.98 g dry weight per g substrate, when supplemented to batch cultures in a glucose-mineral medium. Maximum effect occurred at pH 6.8, at a glycine concentration of 6-12 mmol l(-1), and at cell densities below 1.15 g dry weight l(-1) (0D(610).3). When glycine was added to a culture at a cell density of 1.15 g l(-1) or above, no growth promoting effect of glycine was seen. The 'glycine effect' was not due to CO(2) produced by the glycine cleavage system (GCV), and the lack of effect at higher cell densities was not masked by acetate accumulation, but coincided with increased acetate production. The metabolism of glycine was further investigated in cultures supplied with [2-(13)C] labelled glycine, and the redistribution of label in the [1-(13)C], [2-(13)C], and [1,2-(13)C] isotopomeres of excreted acetate was analysed by 13C NMR. The NMR data revealed that very little degradation of glycine occurred at cell densities below 1.15 g l(-1). Simultaneously the biosynthesis of serine and glycine was repressed as judged by the absence of [2-(13)C] acetate, implying that added glycine was used as a source of glycine, serine, one-carbon units, and threonine. At cell densities above 1.15 g l(-1), 53% of the consumed glycine carbon was excreted as acetate. Degradation of glycine was associated with an increased uptake rate, cleavage by GCV, and degradation of both glycine-derived serine, and glucose-derived serine to pyruvate. This switch in metabolism appears to be regulated by quorum sensing.  相似文献   

7.
Hydrogen cyanide (HCN) production by Pseudomonas aeruginosa in a synthetic medium is stimulated by the presence of glycine. Methionine enhances this stimulation but will not substitute for glycine as a stimulator of cyanogenesis. Threonine and phenylalanine are effective substitutes for glycine in the stimulation of HCN production. Glycine, threonine, and serine are good radioisotope precursors of HCN, but methionine and phenylalanine are not. Cell extracts of P. aeruginosa convert [14C]threonine to [14C]glycine. H14CN is produced with low dilution of label from either [1-14C]glycine or [2-14C]glycine, indicating a randomization of label either in the primary or secondary metabolism of glycine. When whole cells were fed [1,2-14C]glycine, cyanide and bicarbonate were the only radioactive extracellular products observed.  相似文献   

8.
The threonine dehydrogenase (TDG) pathway is a significant route of threonine degradation, yielding glycine in experimental animals, but has not been accurately quantitated in humans. Therefore, the effect of a large excess of dietary threonine, given either as free amino acid (+Thr) or as a constituent of protein (+P-Thr), on threonine catabolism to CO(2) and to glycine was studied in six healthy adult males using a 4-h constant infusion of L-[1-(13)C]threonine and [(15)N]glycine. Gas chromatography-combustion isotope ratio mass spectrometry was used to determine [(13)C]glycine produced from labeled threonine. Threonine intakes were higher on +Thr and +P-Thr diets compared with control (126, 126, and 50 micromol x kg(-1) x h(-1), SD 8, P < 0.0001). Threonine oxidation to CO(2) increased threefold in subjects on +Thr and +P-Thr vs. control (49, 45, and 15 micromol x kg(-1) x h(-1), SD 6, P < 0.0001). Threonine conversion to glycine tended to be higher on +Thr and +P-Thr vs. control (3.5, 3.4, and 1.6 micromol x kg(-1) x h(-1), SD 1.3, P = 0.06). The TDG pathway accounted for only 7-11% of total threonine catabolism and therefore is a minor pathway in the human adult.  相似文献   

9.
To study amino acid exchange between plasma and erythrocytes in vivo, 4-h primed, continuous intravenous infusions of L-[1-13C]leucine, [15N]glycine, and L-[15N]alanine were administered to five healthy young men in the postabsorptive state. Stable isotope enrichments and amino acid levels were determined by gas chromatography-mass spectrometry in both plasma and whole blood and estimated (using hematocrit) in erythrocytes. A high concentration gradient across the erythrocyte membrane was consistently found for glycine (552 +/- 268 microM in erythrocytes vs. 155 +/- 35 microM in plasma), but not for leucine or alanine. A steady-state isotopic enrichment was observed in whole blood as well as plasma for each amino acid in every subject. Steady-state [13C]leucine enrichment in erythrocytes did not differ from plasma enrichment at steady state, the ratio of erythrocyte to plasma enrichment being 1.03 +/- 0.20 (95% confidence limits = 0.78-1.28); in contrast, this ratio reached only 0.23 +/- 0.04 and 0.59 +/- 0.09 (confidence limits 0.18-0.28 and 0.48-0.70) for [15N]glycine and [15N]alanine at steady state, respectively. These results suggest that most of erythrocyte leucine is exchangeable with plasma, whereas only a fraction of erythrocyte glycine and alanine is involved in exchange with plasma in vivo.  相似文献   

10.
The goal of this field study was to provide insight into three distinct populations of microorganisms involved in in situ metabolism of phenol. Our approach measured 13CO2 respired from [13C]phenol and stable isotope probing (SIP) of soil DNA at an agricultural field site. Traditionally, SIP-based investigations have been subject to the uncertainties posed by carbon cross-feeding. By altering our field-based, substrate-dosing methodologies, experiments were designed to look beyond primary degraders to detect trophically related populations in the food chain. Using gas chromatography-mass spectrometry (GC/MS), it was shown that (13)C-labeled biomass, derived from primary phenol degraders in soil, was a suitable growth substrate for other members of the soil microbial community. Next, three dosing regimes were designed to examine active members of the microbial community involved in phenol metabolism in situ: (i) 1 dose of [13C]phenol, (ii) 11 daily doses of unlabeled phenol followed by 1 dose of [13C]phenol, and (iii) 12 daily doses of [13C]phenol. GC/MS analysis demonstrated that prior exposure to phenol boosted 13CO2 evolution by a factor of 10. Furthermore, imaging of 13C-treated soil using secondary ion mass spectrometry (SIMS) verified that individual bacteria incorporated 13C into their biomass. PCR amplification and 16S rRNA gene sequencing of 13C-labeled soil DNA from the 3 dosing regimes revealed three distinct clone libraries: (i) unenriched, primary phenol degraders were most diverse, consisting of alpha-, beta-, and gamma-proteobacteria and high-G+C-content gram-positive bacteria, (ii) enriched primary phenol degraders were dominated by members of the genera Kocuria and Staphylococcus, and (iii) trophically related (carbon cross-feeders) were dominated by members of the genus Pseudomonas. These data show that SIP has the potential to document population shifts caused by substrate preexposure and to follow the flow of carbon through terrestrial microbial food chains.  相似文献   

11.
1. A species of Arthrobacter (designated Arthrobacter 9759) was isolated from soil by its ability to grow aerobically on l-threonine as sole source of carbon atoms, nitrogen atoms and energy; the organism also grew well on other sources of carbon atoms including glycine, but no growth was obtainable on aminoacetone or dl-1-aminopropan-2-ol. 2. During growth on threonine, (14)C from l-[U-(14)C]threonine was rapidly incorporated into glycine and citrate, and thereafter into serine, alanine, aspartate and glutamate. 3. With extracts of threonine-grown cells supplied with l-[U-(14)C]threonine, evidence was obtained of the NAD and CoA-dependent catabolism of l-threonine to produce acetyl-CoA plus glycine. Short-term incorporation studies in which [2-(14)C]acetate and [2-(14)C]glycine were supplied (a) to cultures growing on threonine, and (b) to extracts of threonine-grown cells, showed that the acetyl-CoA was metabolized via the tricarboxylic acid cycle and glyoxylate cycle whereas the glycine was converted into pyruvate via the folate-dependent ;serine pathway'. 4. The threonine-grown organism contained ;biosynthetic' threonine dehydratase and a potent NAD-linked l-threonine dehydrogenase but possessed no l-threonine aldolase activity. 5. Evidence was obtained that the acetyl-CoA and glycine produced from l-threonine had their immediate origin in the alpha-amino-beta-oxobutyrate formed by the threonine dehydrogenase; the CoA-dependent cleavage of this compound was catalysed by an alpha-amino-beta-oxobutyrate CoA-ligase, which was identified with ;aminoacetone synthase'. A continuous spectrophotometric assay of this enzyme was developed, and it was found to be inducibly synthesized only during growth on threonine and not during growth on acetate plus glycine. 6. By using a reconstituted mixture of separately purified l-threonine dehydrogenase and alpha-amino-beta-oxobutyrate CoA-ligase (i.e. ;aminoacetone synthase'), l-[U-(14)C]threonine was broken down to [(14)C]glycine plus [(14)C]acetyl-CoA (trapped as [(14)C]citrate). 7. There was no evidence of aminoacetone metabolism by Arthrobacter 9759 even though a small amount of this amino ketone appeared in the culture medium during growth on threonine.  相似文献   

12.
The native hormones from tassels of maize (Zea mays) were re-investigated. The previous identification by GC/SIM of GA1, GA8 and GA29 in normal tassels was confirmed by full GC/MS scans at the correct Kovats retention indices. In tassels of dwarf-1 mutants, GA44,?GA19, GA17, GA20 and the 16,17-dihydro, 7β,16α,17-trihydroxy derivative of ent-kaurenoic acid were identified by GC/MS. Gibberellin A1 was not found in the mutant tassels. [14C]Gibberellin A53 was fed to tassels of the dwarf-5 mutant. In the ethyl acetate-soluble acidic fraction from the feeds, [14C]GA44 was identified by GC/MS; [14C]GA19 and [14C]GA29 were identified by GC/SIM. The GA29 is probably a metabolite of the feeds because the dwarf-5 mutant is known to control the step copalyl pyrophosphate to ent-kaurene in the maize GA-biosynthetic pathway and because GA29 was not identified in a control experiment. The n-butanol fractions obtained from the feeds were shown, by GC/MS, to contain [14C]GA53 after hydrolysis, suggesting that conjugated [14C]GA53 is a major metabolite from GA53 feeds. [17-13C, 17-3H2]Gibberellin A20 was fed to normal, dwarf-1 and dwarf-5 tassels. In each case, analysis of the purified ethyl acetate-soluble acidic extracts by GC/MS led to the identification of [13C]GA29 and unmetabolized [13C]GA20 in which no 13C-isotope dilution was observed.  相似文献   

13.
[1,2,3,4-13C]cortisol was i.v. administered to two sisters aged 11 yr (patient I) and 3 yr (patient II) who suffer from 17 alpha-hydroxylase deficiency. This is the first time that the cortisol production rate (CPR) in patients with 17 alpha-hydroxylase deficiency has been measured with a stable labelled tracer using the urinary method. The urine was collected for 3 days. High-performance liquid chromatography (HPLC) of approximately 100 ml urine extracts was carried out to isolate the small amount of cortisol metabolites excreted. The cortisol metabolites were oxidized to 11-oxo-aetiocholanolone. The isotope dilution in the methyl oxime tert-butyldimethylsilyl ether derivatives was measured by selected ion monitoring gas chromatography/mass spectrometry (GC/MS). The CPR calculated from tetrahydrocortisone (THE) and the cortolones was 765 and 536 nmol/day, respectively in patient I. The CPR in patient II was only calculated from THE and was 62 nmol/day. If radioactive labelled cortisol had been used, much larger quantities of urine would have been needed for isolation of sufficient mass of metabolites, even then purification may have been difficult. Steroid profiling of 1 ml urine samples by GC and identification by GC/MS revealed high concentrations of pregnenolone, progesterone, 11 beta-hydroxy progesterone and corticosterone metabolites. Tetrahydrocorticosterone and 5 alpha-tetrahydrocorticosterone were found in urine at elevated excretions of 2.5 and 5.7, 0.9 and 2.0 mumols/24 h, in patients I and II respectively. No cortisol metabolites were detected by routine GC or GC/MS as the low amounts excreted co-eluted with the relatively abundant corticosterone metabolites.  相似文献   

14.
We describe a simple method using (13)C labeling and NMR spectroscopy to determine the flux contributions of alternative pathways in Saccharomyces cerevisiae that produce the same metabolite with identical labeling patterns. Cells were incubated with a (13)C-labeled precursor for one of the branches, and the absolute enrichment of the product and its metabolic precursor(s) was quantified. The ratio of the absolute enrichment of the product to that of its precursor reflects the contribution of the pathway. The method was applied to the biosynthesis of glycine in yeast, which can occur from threonine via threonine aldolase or from serine via serine hydroxymethyltransferase. [2-(13)C]Aspartate and [2-(13)C]serine were used as labeled precursors for the threonine aldolase and serine hydroxymethyltransferase pathways, respectively. The data show that in cells possessing both pathways, the serine hydroxymethyltransferase pathway contributes 65-75% of the total glycine production. In comparison with other approaches, this method provides an inexpensive, flexible alternative to determining the flux contributions of split pathways under controlled conditions and should have wide applicability in the metabolic engineering of microorganisms.  相似文献   

15.
Saccharomyces cerevisiae has both cytoplasmic and mitochondrial C1-tetrahydrofolate (THF) synthases. These trifunctional isozymes are central to single-carbon metabolism and are responsible for interconversion of the THF derivatives in the respective compartments. In the present work, we have used 13C NMR to study folate-mediated single-carbon metabolism in these two compartments, using glycine and serine synthesis as metabolic endpoints. The availability of yeast strains carrying deletions of cytoplasmic and/or mitochondrial C1-THF synthase allows a dissection of the role each compartment plays in this metabolism. When yeast are incubated with [13C]formate, 13C NMR spectra establish that production of [3-13C]serine is dependent on C1-THF synthase and occurs primarily in the cytosol. However, in a strain lacking cytoplasmic C1-THF synthase but possessing the mitochondrial isozyme, [13C]formate can be metabolized to [2-13C]glycine and [3-13C]serine. This provides in vivo evidence for the mitochondrial assimilation of formate, activation and conversion to [13C]CH2-THF via mitochondrial C1-THF synthase, and subsequent glycine synthesis via reversal of the glycine cleavage system. Additional supporting evidence of reversibility of GCV in vivo is the production of [2-13C]glycine and [2,3-13C]serine in yeast strains grown with [3-13C]serine. This metabolism is independent of C1-THF synthase since these products were observed in strains lacking both the cytoplasmic and mitochondrial isozymes. These results suggest that when formate is the one-carbon donor, assimilation is primarily cytoplasmic, whereas when serine serves as one-carbon donor, considerable metabolism occurs via mitochondrial pathways.  相似文献   

16.
The stereochemical course of the vitamin K dependent carboxylation has been elucidated using a (4S)-4-fluoroglutamyl-containing pentapeptide as a substrate. The absolute configuration of the [13C]-4-carboxy-4-fluoroglutamate obtained when the carboxylation was carried out with 13C-labeled sodium bicarbonate, was determined after reduction of the [13C]-4-carboxy-4-fluoroglutamyl residue into 4-fluoro-5,5'-dihydroxyleucine, hydrolysis, lactonization, and peracetylation. The absolute configuration at C-4 was determined to be S by locating the 13C label in the lactone ring of the trans isomeric lactone and in the hydroxymethyl group of the cis isomer following HPLC separation of both isomers and analysis by GC/MS/MS techniques. It follows that the vitamin K dependent carboxylation occurs with inversion of configuration.  相似文献   

17.
The molecular dynamics and structural organization of mouse epidermal keratin intermediate filaments (IF) have been studied via solid-state nuclear magnetic resonance (NMR) experiments performed on IF labeled both in vivo and in vitro with isotopically enriched amino acids. As a probe of the organization of the peripheral glycine-rich end domains of the IF, carbon-13 NMR experiments have been performed on subfilamentous forms (prekeratin) and on IF reassembled in vitro that had been labeled with either [1-13C]glycine or [2-13C]glycine, as more than 90% of the glycines of the keratins are located in the end domains. Although cross-labeling to seryl residues was observed, the proportion of serine located in the end domains is nearly the same as that for glycine. Measurements of carbon relaxation times, nuclear Overhauser enhancements, and signal intensities show that the motions of the peptide backbone in the end domains are effectively isotropic, with average correlation times distributed over the range of 0.2-20 ns. These results indicate that the end domains of IF are remarkably flexible and have little or no structural order. To probe the structural organization of the coiled-coil rod domains of the IF, separate samples of native keratin IF, raised in primary tissue culture, were labeled with L-[1-13C]leucine, L-[2H10]leucine, or L-[2,3,3-2H3]leucine, as greater than 90% of the leucyl residues of the keratin IF types studied are located in the coiled coils which form the central core of IF.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
The current study was undertaken to gain further insight into the carbohydrate metabolism of the proliferating larvae of Mesocestoides corti by studying the effect of the metabolic inhibitors oxaloacetate and iodoacetate on the incorporation of label from NaH13CO3 into the excretory products lactate and succinate. These products were converted to their methyl derivatives and analyzed by gas chromatography/mass spectrometry (GC/MS) to determine the isotope ratios. The percentage of [1-13C]lactate, [1-13C]succinate, and [1,4-13C2]succinate decreased in the presence of oxaloacetate; iodoacetate did not significantly influence the relative amount of the former, but markedly increased percentages of the latter two. The data provide information on the relative rates of the pathways and steps within them and the influence of the inhibitors on these pathways.  相似文献   

19.
Viramidine is a prodrug of ribavirin. To facilitate pharmacokinetics studies of viramidine in man, a sensitive and specific LC-MS/MS method for the simultaneous analyses of viramidine and ribavirin in human plasma was developed and validated. The method involved the addition of [13C]viramidine and [13C]ribavirin as internal standards, protein precipitation with acetonitrile, HPLC separation, and quantification by MS/MS system using positive electrospray ionization in the multiple reaction monitoring mode (MRM). The precursor-->product ion transitions were monitored at 245-->113, 250-->113, 244-->112, and 249-->112 for ribavirin, [13C]ribavirin, viramidine, and [13C]viramidine, respectively. The calibration curves for viramidine and ribavirin were linear over a concentration range of 1-1000 ng/mL. For both viramidine and ribavirin, the lower limit of quantification (LLOQ) was 1 ng/mL. For viramidine, intra- and inter-day analyses of QC samples at 1, 5, 250, and 1000 ng/mL indicated good precision (%CV between 1.0 and 7.0%) and accuracy (%bias between -4.3 and 5.2%). For ribavirin, intra- and inter-day analyses of QC samples at 1, 5, 250, and 1000 ng/mL indicated similar precision (%CV between 0.8 and 8.3%) and accuracy (%bias between -5.8 and 9.4%). Both viramidine and ribavirin were stable in human plasma stored at room temperature for at least 3 h, 4 degrees C for at least 6 h, and for at least three freeze-thaw cycles. This accurate and highly specific assay provides a useful method for evaluating the pharmacokinetics of viramidine and ribavirin in man following administration of viramidine.  相似文献   

20.
Loik ME  Nobel PS 《Plant physiology》1993,103(3):871-876
The responses to low temperature were determined for two species of cacti sensitive to freezing, Ferocactus viridescens and Opuntia ficus-indica, and a cold hardy species, Opuntia fragilis. Fourteen days after shifting the plants from day/night air temperatures of 30/20[deg]C to 10/0[deg]C, the chlorenchyma water content decreased only for O. fragilis. This temperature shift caused the freezing tolerance (measured by vital stain uptake) of chlorenchyma cells to be enhanced only by about 2.0[deg]C for F. viridescens and O. ficus-indica but by 14.6[deg]C for O. fragilis. Also, maintenance of high water content by injection of water into plants at 10/0[deg]C reversed the acclimation. The endogenous abscisic acid (ABA) concentration was below 0.4 pmol g-1 fresh weight at 30/20[deg]C, but after 14 d at 10/0[deg]C it increased to 84 pmol g-1 fresh weight for O. ficus-indica and to 49 pmol g-1 fresh weight for O. fragilis. Four days after plants were sprayed with 7.5 x 10-5 M ABA at 30/20[deg]C, freezing tolerance was enhanced by 0.5[deg]C for F. viridescens, 4.1[deg]C for O. ficus-indica, and 23.4[deg]C for O. fragilis. Moreover, the time course for the change in freezing tolerance over 14 d was similar for plants shifted to low temperatures as for plants treated with exogenous ABA at moderate temperatures. Decreases in plant water content and increases in ABA concentration may be important for low-temperature acclimation by cacti, especially O. fragilis, which is widely distributed in Canada and the United States.  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