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 20 毫秒
1.
Using microdialysis, we compared intracerebral and subcutaneous administration of nicotine for the effect on the levels of extracellular amino acids in the hippocampus of anesthetized rats. Administration by microdialysis of 10 mM nicotine, resulting in a nicotine concentration of 0.134 μmol/g in the hippocampus, increased the extracellular levels of aspartic acid, glutamic acid, and serine by 26–60%. At 50 mM nicotine the increases in the levels of aspartic acid, glutamic acid, serine, glycine, and glutamine were between 76% and 141%. Subcutaneous administration of nicotine at a dose of 6 μmol/kg caused a 57% increase in the extracellular level of glutamic acid. After a dose of 12 μmol/kg that resulted in a nicotine level of 0.015 μmol/kg in the hippocampus, the extracellular level of glutamic acid was increased by 100%, and that of aspartic acid by 24%. Thus, higher cerebral nicotine levels were needed with intracerebral than with subcutaneous administration to obtain similar amino acid changes. Prior administration of mecamylamine or L-kynurenine prevented the subcutaneous nicotine-induced elevation of the extracellular levels of aspartic acid and glutamic acid. Our results indicate that receptor interactions modulate nicotine effects and that both nicotinic cholinergic and NMDA/glycine glutamatergic receptors participate in the action of nicotine in increasing extracellular amino acid levels.  相似文献   

2.
The amino acids lysine and glycine are reported to react with glucose at physiological pH and temperature and undergo non-enzymic glycation. Three other amino acids present in relatively larger amounts in the lens i.e. alanine, aspartic acid and glutamic acid were also found to undergo non-enzymic glycation as found by incorporation of uniformly labelled (U-[14C]) glucose into the amino acids. The glucose incorporation was 1.6 to 2.5% for alanine, 35 to 50% for aspartic acid and 2.3 to 3.3% for glutamic acid. Each amino acid of varying concentrations lowered the extent ofin vitro glycation of lens proteins significantly in glucose-treated homogenates of normal lens from humans. The decrease in glycation for alanine was between 32 and 69%, that for aspartate was between 18 and 74%, and for glutamate was between 52 to 74%. Decreased glycation was greater for higher concentrations of glucose. Scavenging of intracellular glucose and decreasing the extent of glycation of lens proteins could be the mechanism of action by which the amino acids alanine, aspartic acid and glutamic acid could exercise a beneficial effect on cataract and diabetic retinopathy.  相似文献   

3.
4.
Abstract

Qualitative and quantitative analysis of free and bound amino acids and amides during dormancy and the most important phases of the first cell cycle was carried out in tubers of Helianthus tuberosus.

In the dormant tuber arginine was confirmed to be the most abundant amino acid. A high amount of asparagine was also present; on the contrary glutamine was found in very low concentrations. During the progression of dormancy, all the free amino acids and amides declined while aspartic and glutamic acid increased.

During the G1 phase of the first cell cycle induced by 2,4-D, all the free amino acids and amides decreased with the exception of glutamic acid.

At 18, 20, 24 h of activation with 2,4-D, corresponding to the S phase and the beginning of mitosis, bound amino acids were also determined. In these phases of the cell cycle they increased reaching a maximum at 20 h; on the other hand the free amino acid and amide content, especially aspartic acid, asparagine and arginine, decreased with the exception of glutamic acid, alanine and phenylalanine.  相似文献   

5.
We present an analysis of intron positions in relation to nucleotides, amino acid residues, and protein secondary structure. Previous work has shown that intron sites in proteins are not randomly distributed with respect to secondary structures. Here we show that this preference can be almost totally explained by the nucleotide bias of splice site machinery, and may well not relate to protein stability or conformation at all. Each intron phase is preferentially associated with its own set of residues: phase 0 introns with lysine, glutamine, and glutamic acid before the intron, and valine after; phase 1 introns with glycine, alanine, valine, aspartic acid, and glutamic acid; and phase 2 introns with arginine, serine, lysine, and tryptophan. These preferences can be explained principally on the basis of nucleotide bias at intron locations, which is in accordance with previous literature. Although this work does not prove that introns are inserted into genomes at specific proto-splice sites, it shows that the nucleotide bias surrounding introns, however it originally occurred, explains the observed correlations between introns and protein secondary structure.  相似文献   

6.
Effect of L-Aspartic Acid and L-Glutamic Acid on Production of L-Proline   总被引:2,自引:1,他引:1  
To elucidate the effect of aspartic acid on growth of Kurthia catenaforma during the proline fermentation, this organism was compared with other bacteria with respect to the rate of consumption of aspartic acid, and to the activities of enzymes concerned in the metabolism of aspartic acid. Although no marked difference in enzyme activities was observed, the aspartic acid consumption rate of K. catenaforma was markedly higher than that of other organisms. The consumption of glutamic acid by K. catenaforma was not detected at 24 hr of culture. The difference between the consumption of aspartic acid and glutamic acid in this strain might result from a difference in permeability to the amino acids. We considered that L-glutamic acid might substitute for L-aspartic acid if the uptake of glutamic acid could be increased. A number of detergents were screened for their effect on consumption of glutamic acid. Cetyltrimethylammonium bromide, sodium laurylphosphate, and polyoxyethylene sorbitan monolaurate were found to increase the transport rate of glutamic acid, but not of aspartic acid. A method of producing L-proline from glutamic acid was established with the aid of detergents.  相似文献   

7.
Domoic acid is a shellfish toxin which produces gastrointestinal distress, followed by neurological symptoms such as headache, confusion, disorientation and severe deficits in short-term memory. Domoic acid is an amino acid which contains three carboxylic groups, and one imino group, and its solubility, rate of absorption, and elimination would vary depending on the protonation of these groups at different pH's. We propose that domoic acid toxicity varies with pH of administered domoic acid solution. Domoic acid toxicity was measured in mice as the onset times for scratching behaviour, seizure activity, and death, after the intraperitoneal administration of domoic acid at different pH's. Results of the present study show that the scratching behaviour, seizure activity, and death, occurred at 12, 40, and 55 min, after intraperitoneal administration of domoic acid at pH 3.7. Apparently, the onset times for three types of behaviours were relatively long, and well separated from each ot her. Domoic acid toxicity was lowest at pH 3.7, and highest at pH 7.4, with intermediate toxicity at other pH's. The onset time of scratching behaviour was not influenced by pH of domoic acid solution at three different doses. In contrast, the onset times for seizure activity, and death were significantly affected by pH of domoic acid, toxicity being higher at pH 7.4 than at pH 3.7. The pH effect on domoic acid toxicity diminished as the dose of domoic acid was increased. In fact, at 14.5 mg/kg domoic acid toxicity was similar at both pH's of 3.7 and 7.4. It is concluded that in vivo toxicity of domoic acid varies depending on pH of the administered solution. The differential toxicity of domoic acid at different pH may be related to its solubility, rate of absorption, and elimination, depending on the degree of protonation of domoic acid molecule. Domoic acid toxicity would also vary depending on the age of animal, receptor sensitivity and density in different regions of brain. (Mol Ce ll Biochem 167: 179-185, 1997)  相似文献   

8.
Metabolites of -[14C]proline were found in the trichloroacetic acid-soluble fraction of 16-day-old chick embryo frontal bones. In several ion-exchange procedures these metabolites interfered with the analysis of hydroxyproline derived from the metabolic breakdown of collagen. The major metabolite was identified as glutamic acid by its chromatographic and crystallization properties. It was eluted from AG50 cation-exchange resin with 1.0 HCL in the hydroxyproline region, but was separated from hydroxyproline on a DC-6A column in the amino acid analyzer. Another metabolite was identified as aspartic acid. It was not separated from hydroxyproline on either AG50 using 1 HCL for elution or on DC-6A using 0.1 sodium citrate, pH 3.25, for elution, but adequate separation was obtained by elution with 0.2 sodium citrate buffer at pH 2.91. Formation of these metabolites was not related either to protein synthesis or proline hydroxylation. Therefore, it is possible to analyze for hydroxyproline accurately by using a separate unhydroxylated sample to correct for the presence of the metabolites. The formation of glutamic acid suggested that proline oxidase activity might be present in bone tissue, but none was detected using a sensitive radioisotopic assay. Although the amount of radioactivity found in the metabolites was 36% of the amount of [14C]proline incorporated into protein, no radioactive glutamic or aspartic acid was present in protein hydrolyzates. This observation suggests that the metabolites did not enter the major amino acid pool used for protein synthesis.  相似文献   

9.
Domoic acid, a recognized neurotoxin derived from contaminated samples of the blue mussel (Mytilus edulis L.), was analyzed for mutagenicity at 2 loci and for 2 cytogenetic parameters in a hepatocyte-mediated assay with V79 Chinese hamster lung fibroblasts. Genetic end-points measured were: mutation to 6-thioguanine resistance at the HGPRTase locus; mutation to ouabain resistance at the Na+,K+-ATPase locus; sister-chromatid exchange (SCE) and micronucleus frequency (MN). None of these genetic end-points was significantly affected by exposure to domoic acid at dose levels of 27.2 and 54.4 micrograms/ml with or without activation by freshly isolated rat liver hepatocytes. It was concluded that, within the limits of the test system employed, domoic acid was non-genotoxic to V79 cells.  相似文献   

10.
The fate of aspartic acid used for proline fermentation by Kurthia catenaforma was traced by using aspartic acid-U-(14)C. The radioactivities of proline and glutamic acid increased with the disappearance of aspartic acid. After 40 hr, aspartic acid disappeared from the medium and radioactive alpha-ketoglutaric acid was detected. The radioactivity of proline reached 44% of aspartic acid radioactivity at 40 hr. The specific radioactivities of these amino acids and of alpha-ketoglutaric acid supported the notion that proline is produced mainly from aspartic acid via alpha-ketoglutaric acid and glutamic acid. Since the levels of glutamic acid dehydrogenases (EC 1.4.1.2 and EC 1.4.1.4) were low in this organism, it appears that the nitrogen atom of aspartic acid enters proline by the action of aspartate aminotransferase (EC 2.6.1.1). The mechanism of proline production is discussed on the basis of the role of aspartic acid in this fermentation.  相似文献   

11.
Domoic acid in phytoplankton and fish in San Diego, CA, USA   总被引:1,自引:0,他引:1  
We provide the first confirmation of the presence of domoic acid (DA) in phytoplankton and fish in San Diego, California, based on samples collected between 1 October 2003 and 29 September 2004. In February 2004, we detected DA in seawater samples collected off the Scripps Pier and also in coastal samples as far as 120 km to the north. At the same time we observed populations of toxic Pseudo-nitzschia australis and Pseudo-nitzschia multiseries as high as 7.7 × 104 cells l−1. Elevated concentrations of DA and abundances of the toxic species were also found further north in coastal waters of Orange County and, to a lesser extent, in southern Los Angeles County. DA concentrations in the viscera from four species of fish obtained at or near the Scripps Pier ranged from low to above the critical level for public safety. Samples of mussel tissues from the Scripps Pier analyzed by the State Department of Health Services contained low but detectable amounts of DA. Concomitant sea lion strandings from San Diego to Malibu Beach may be related to the presence of DA. DA in tissue from mussels and fish provides evidence for the local transfer of DA from an algal source to higher trophic levels in San Diego coastal waters.  相似文献   

12.
Abstract— The influence of hypercapnia on the metabolism of glutamic acid, aspartic acid, glutamine and GABA in rat brain was studied using three different precursors. Acute hypercapnia induced a fall in the concentration of glutamic and aspartic acid, and a rise in the concentration of glutamine and GABA. Acute hypercapnia had a profound effect on the relative specific radioactivity of glutamine indicating that the excess glutamine, present in the brain in hypercapnia, was synthetized from glutamic acid in the compartment where it could become quickly labelled from butyric and glutamic acid, but not from glucose. This effect was maintained in chronic hypercapnia.  相似文献   

13.
Montuelle  B.  Kestemont  P.  Chalamet  A. 《Hydrobiologia》1992,243(1):71-77
Bacterial degradation of a high concentration of amino acids (up to 150 mg l–1), released by quail manure pellets, was studied in a fertilized fish pond. Aspartic acid, Glutamic acid and Alanine were major components of the free amino acid pool (respectively 14.6%, 16.4% and 11.1%). Laboratory experiments showed a fast decomposition of these three amino acids. The degradation was a zero order reaction. Activation energy, Ea, was of the same order for the three amino acids (from 64.2 to 70.3 kg-mol–1). There was no significant difference between the behaviour of aspartic acid in the presence or not of other carbon and nitrogen supply, but stimulation of degradation of glutamic acid and of Alanine by NH4 + and C, was marked. This process may be important in the initial mineralisation of the fertilizer in fish ponds.  相似文献   

14.
The amplified expression of a recombinant protein is known to lead to an intracellular depletion of specific amino acid pools which in turn may affect the production of the desired protein. In order to counteract and overcome such a situation during the fermentation of the recombinant Escherichia coli (PMSG27) containing the glucose isomerase (GI) gene from Streptomyces sp. NCIM 2730, the effect of addition of different amino acids on the specific activity of GI was studied. The amino acid composition of GI from Streptomyces sp. NCIM 2730 reveals predominantly aspartic acid, glutamic acid, and glycine; therefore, in the present paper, the effect of coordinated addition of the assorted combinations of these three amino acids on the synthesis of recombinant GI was studied. The results were analyzed using a 23 factorial design. The following conclusions were derived from the analysis of two-factor interactions of the three amino acids: (i) The interaction between the aspartic and glutamic acid is independent of aspartic acid concentration but is affected by the increasing concentrations of glutamic acid, (ii) The effect of aspartic acid concentration is more than that of glycine, and (iii) During the interaction of glutamic acid and glycine, the effect of glutamic acid is more prominent than that of glycine. The three-factor interaction analyses reveal that the effect of the three amino acids is in the order aspartic acid > glutamic acid > glycine.  相似文献   

15.
The accumulation of certain essential and metabolically derived amino acids in the free amino acid pools of three excitable tissues has been studied in the chick embryo. Valine together with leucine are at the onset present in the yolk at higher concentrations than any of the other essential amino acids. By 15 days all the amino acids studied have accumulated in the embryonic pools at a higher rate than valine, although certain amino acids, such as phenylalanine or methionine, always remain at lower relative concentrations than valine. This reflects their low supply in the yolk, rather than a more rapid rate of disappearance (utilization). During early embryogenesis (E2–E4), tissues preferentially concentrate glutamic acid, besides taurine and phosphoethanolamine (6). The next distinct stage of development (E4–E7) is marked in the brain by a gradual rise in glutamic acid, glutamine and aspartic acid; the same three amino, acids do not demonstrate a further increase in the pool of the heart, while in the whole eye the amino acid profile begins to resemble the blood. Leucine in all three tissues declines rapidly, to reach isoleucine levels by day 7 of development; tyrosine increases slowly in apparent reciprocity to an equally gradual phenylalanine decrease. Into the second week of embryo growth (E7–E15), GABA appears in the mesencephalon (E7) and the eye (E9–E10). In the mesencephalon, the free amino acid pool composition exhibits a rather sudden increase of most metabolically derived amino acids. Glutamic acid and glutamine in the brain increase in parallel; the rate of GABA and aspartic acid accumulation is slower, and for GABA stabilizes on day 14, as does glutamine. In the eye, by day 15, GABA levels are more closely aligned with the aspartic acid content. Finally, throughout embryogenesis serine fluctuations in blood and tissues are in parallel with those of threonine, and different from glycine or alanine which also change in tandem.  相似文献   

16.

Settlement of barnacle larvae is believed to be induced by the chemical cues present in their surrounding environment. Here, an investigation was carried out on the effects of sixteen different mono‐amino acids with acidic, basic, uncharged polar and nonpolar side chains, and GABA on larval settlement of the barnacle, Balanus amphitrite. Settlement inducing activity by nine mono‐amino acids, viz. asparagine, glutamine, tyrosine, serine, glycine, tryptophan, leucine, isoleucine and valine (but not phenylalanine) with uncharged polar and nonpolar side chains was observed. Of these, the most active mono‐amino acids were serine, leucine and isoleucine, which were effective at a threshhold of 1.0 × 10‐7 M. On the other hand, aspartic acid, glutamic acid, GABA, and the basic mono‐amino acids lysine, arginine and histidine did not have any inducing effect. These results suggest that uncharged polar and non‐polar end group of the amino acid chain play an important role in inducing the settlement process in cyprids.  相似文献   

17.
Increased plasma glutamic acid in a genetic model of epilepsy   总被引:4,自引:0,他引:4  
A significant increase in the plasma levels of glutamic acid and a significant decrease in aspartic acid and taurine in epileptic patients and their first degree relatives was reported more than a decade ago and an underlying genetic basis for these amino acid changes was suggested. The main objective of the present study was to determine the plasma levels of glutamic acid, aspartic acid and taurine in El mice which are an inbred epileptic mutant mouse strain. The results show a significant increase in plasma glutamic acid but no changes in aspartic acid or taurine in the epileptic mice as compared to controls. The data provide the first evidence of a significant increase in plasma glutamic acid in an animal model of hereditary epilepsy and substantiate the hypothesis that a genetic defect underlies the elevated plasma glutamic acid levels in association with epilepsy. The findings are also compatible with neurochemical and neurophysiological evidence implicating glutamic acid in the mechanism of seizures.  相似文献   

18.
The presence of aluminium in amino acids parenteral nutrition solutions can be related to the affinity of the amino acids for aluminium present in glass containers used for storage. For this study solutions of 19 amino acids used in parenteral nutrition were stored individually in glass flasks and the aluminium measured at determined time intervals. Solutions of complexing agents for aluminium, as ethylene-diaminetetraacetic acid, nitrilotriacetic acid, citrate, oxalate and fluoride ions were also stored in the same flasks and the aluminium measured during the same time interval. The measurements were made by electrothermal atomic absorption spectrometry. The aluminium content of the glass containers was also measured. The results showed that the glasses have from 0.6% to 0.8% Al. Only solutions of cysteine, cystine, aspartic acid and glutamic acid became contaminated by aluminium. As the same occurred with the complexing agents, aluminum can be released from glass due to an affinity of the substances for aluminium. Comparing the action of complexing agents and amino acids for which the stability constants of aluminium complex are known, it is possible to relate the magnitude of the stability constant with the aluminium leached from glass, the higher the stability constant, the higher the aluminium released. The analysis of commercial formulations with and without cysteine, cystine, glutamic acid or aspartic acid stored in glass containers confirms that the presence of these amino acids combined with the age of the soLution are, at least partially, responsible for the aluminium contamination. The resuLts demonstrated that the contamination is an ongoing process due to the presence of aluminium in glass combined with the affinity of some amino acids for this element.  相似文献   

19.
Gas chromatography-mass spectrometry studies of the nitrogen isotopic composition of the N-trifluoroacetyl n-butyl ester derivatives of the amino acids from isolated hydrolyzed cyanophycin from 15N-enriched cells led to two major findings: (1) the amino acid composition of this granular polypeptide, isolated using procedures optimized for extracting and purifying cyanophycin from cells in the stationary growth phase, varied with the culture growth condition; (2) the rate of incorporation of exogenous nitrate differed for each nitrogen atom of the amino acid constituents of cyanophycin or cyanophycin-like polypeptide. Arginine and aspartic acid were the principle components of cyanophycin isolated from exponentially growing cells and from light-limited stationary phase cells, with glutamic acid as an additional minor component. The cyanophycin-like polypeptide from nitrogen-limited cells contained only aspartic and glutamic acids, but no arginine. The glutamic acid content decreased and arginine content increased as nitrate was provided to nitrogen-limited cells. These cells rapidly incorporated nitrate at different rates at each cyanophycin nitrogen site: guanidino nitrogens of arginine>aspartic acid >-amino nitrogen of arginine>glutamic acid. Little media-derived nitrogen was incorporated into cyanophycin of exponentially growing cells during one cellular doubling time.Abbreviations asp-TAB, glu-TAB, arg-TAB N-Trifluoroacetyl n-butyl ester derivatives of aspartic acid, glutamic acid and arginine, respectively - CAP chloramphenicol - CF correction factor - TFAA Trifluoroacetic anhydride - MBTFA N-Methyl-bis-trifluoroacetamide  相似文献   

20.
Domoic acid, an excitatory amino acid structurally related to kainate, was recently identified as being presumably responsible for the recent severe intoxication presented by more than 100 people having eaten mussels grown in Prince Edward Island (Canada). The amino acid kainate has been shown to be highly neurotoxic to the hippocampus, which is the most sensitive structure in the central nervous system. The present in vivo electrophysiological studies were undertaken to determine if domoic acid exerts its neurotoxic effect via kainate receptor activation. Unitary extracellular recordings were obtained from pyramidal neurons of the CA1 and the CA3 regions of the rat dorsal hippocampus. The excitatory effect of domoic acid applied by microiontophoresis was compared with that of agonists of the three subtypes of glutamatergic receptors: kainate, quisqualate, and N-methyl-D-aspartate. In CA1, the activation induced by domoic acid was about threefold greater than that induced by kainate; identical concentrations and similar currents were used. In CA3, domoic acid was also three times more potent than kainate. However, the most striking finding was that domoic acid, similar to kainate, was more than 20-fold more potent in the CA3 than in the CA1 region, whereas no such regional difference could be detected with quisqualate and N-methyl-D-aspartate. As the differential regional response of CA1 and CA3 pyramidal neurons to kainate is attributable to the extremely high density of kainate receptors in the CA3 region, these results provide the first electrophysiological evidence that domoic acid may produce its neurotoxic effects through kainate receptor activation.  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