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.
2.
To explore the effect of sodium l-aspartate monohydrate (aspartate) as a NAD+ regenerating agent for acetaldehyde in alcohol metabolism, a simple HPLC method has been developed for the measurement of free acetaldehyde in total mice blood digested with alcohol and aspartate. The blood samples were collected in EDTA Vacutainer tubes, and treated with 2,4-dinitrophenylhydrazine (DNP hydrazine) reagent in total blood. Acetaldehyde DNP hydrazone was extracted from total blood and analyzed by HPLC using an Ultrasphere ODS column. The compounds were separated using acetonitrile–water (50:50, v/v) as mobile phase and detected at 356 nm. The detection limit for acetaldehyde DNP hydrazone was 0.1 ppm. A blank determination was carried out for each analysis and subtracted from the results. The amount of acetaldehyde in blood has been determined as a function of time lapse after sole alcohol administration and aspartate ingestion followed by alcohol administration, respectively. This comparative analysis demonstrates that the ingestion of aspartate before the administration of alcohol dramatically decreases the aldehyde level in blood, and aspartate may be utilized as a prospective antagonist for acceleration of ethanol metabolism and prevention of acetaldehyde toxicity.  相似文献   

3.
Plant monofunctional aspartate kinase is unique among all aspartate kinases, showing synergistic inhibition by lysine and S-adenosyl-l-methionine (SAM). The Arabidopsis genome contains three genes for monofunctional aspartate kinases. We show that aspartate kinase 2 and aspartate kinase 3 are inhibited only by lysine, and that aspartate kinase 1 is inhibited in a synergistic manner by lysine and SAM. In the absence of SAM, aspartate kinase 1 displayed low apparent affinity for lysine compared to aspartate kinase 2 and aspartate kinase 3. In the presence of SAM, the apparent affinity of aspartate kinase 1 for lysine increased considerably, with K(0.5) values for lysine inhibition similar to those of aspartate kinase 2 and aspartate kinase 3. For all three enzymes, the inhibition resulted from an increase in the apparent K(m) values for the substrates ATP and aspartate. The mechanism of aspartate kinase 1 synergistic inhibition was characterized. Inhibition by lysine alone was fast, whereas synergistic inhibition by lysine plus SAM was very slow. SAM by itself had no effect on the enzyme activity, in accordance with equilibrium binding analyses indicating that SAM binding to aspartate kinase 1 requires prior binding of lysine. The three-dimensional structure of the aspartate kinase 1-Lys-SAM complex has been solved [Mas-Droux C, Curien G, Robert-Genthon M, Laurencin M, Ferrer JL & Dumas R (2006) Plant Cell18, 1681-1692]. Taken together, the data suggest that, upon binding to the inactive aspartate kinase 1-Lys complex, SAM promotes a slow conformational transition leading to formation of a stable aspartate kinase 1-Lys-SAM complex. The increase in aspartate kinase 1 apparent affinity for lysine in the presence of SAM thus results from the displacement of the unfavorable equilibrium between aspartate kinase 1 and aspartate kinase 1-Lys towards the inactive form.  相似文献   

4.
The protein (Escherichia coli CheY) that controls the direction of flagellar rotation during bacterial chemotaxis has been shown to be phosphorylated on the aspartate 57 residue. The residue phosphorylated is present within a conserved sequence in every member of a family of bacterial regulatory proteins. The phosphorylation is transient, with a much shorter half-life than that expected of a simple acyl phosphate intermediate, indicating that the sequence and conformation of the protein is designed to achieve a rapid hydrolysis. The CheY-phosphate linkage can be reductively cleaved by sodium borohydride. High-performance tandem mass-spectrometric analysis of proteolytic peptides derived from [3H]borohydride-reduced phosphorylated CheY protein was used to identify the position of phosphorylation. Mutants with altered aspartate 57 exhibited no chemotaxis. When aspartate 13, another conserved residue, was changed, greatly reduced chemotaxis was observed, suggesting an important role for aspartate 13. The rate-determining step of chemotactic signaling is governed by the kinetics of formation and hydrolysis of the CheY protein phosphoaspartate bond. The CheY protein apparently functions as a protein phosphatase that possesses a transient covalent intermediate. Transient phosphorylation of an aspartate residue is an effective mechanism for producing a biochemical signal with a short concentration-independent half-life. The duration of the signal can be controlled by small structural elements within the phosphorylated protein.  相似文献   

5.
Aspartate transcarbamylase from Bacillus subtilis has been purified to apparent homogeneity. A subunit molecular weight of 33,500 +/- 1,000 was obtained from electrophoresis in polyarcylamide gels containing sodium dodecyl sulfate and from sedimentation equilibrium analysis of the protein dissolved in 6 M guanidine hydrochloride. The molecular weight of the native enzyme was determined to be 102,000 +/- 2,000 by sedimentation velocity and sedimentation equilibrium analysis. Aspartate transcarbamylase thus appears to be a trimeric protein; cross-linking with dimethyl suberimidate and electrophoretic analysis confirmed this structure. B. subtilis aspartate transcarbamylase has an amino acid composition quite similar to that of the catalytic subunit from Escherichia coli aspartate transcarbamylase; only the content of four amino acids is substantially different. The denaturated enzyme has one free sulfhydryl group. Aspartate transcarbamylase exhibited Michaelis-Menten kinetics and was neither inhibited nor activated by nucleotides. Several anions stimulated activity 2- to 5-fold. Immunochemical studies indicated very little similarity between B. subtilis and E. coli aspartate transcarbamylase or E. coli aspartate transcarbamylase catalytic subunit.  相似文献   

6.
The aspartate kinase (AK) from Mycobacterium tuberculosis (Mtb) catalyzes the biosynthesis of aspartate family amino acids, including lysine, threonine, isoleucine and methionine. We determined the crystal structures of the regulatory subunit of aspartate kinase from Mtb alone (referred to as MtbAKβ) and in complex with threonine (referred to as MtbAKβ-Thr) at resolutions of 2.6 ? and 2.0 ?, respectively. MtbAKβ is composed of two perpendicular non-equivalent ACT domains [aspartate kinase, chorismate mutase, and TyrA (prephenate dehydrogenase)] per monomer. Each ACT domain contains two α helices and four antiparallel β strands. The structure of MtbAKβ shares high similarity with the regulatory subunit of the aspartate kinase from Corynebacterium glutamicum (referred to as CgAKβ), suggesting similar regulatory mechanisms. Biochemical assays in our study showed that MtbAK is inhibited by threonine. Based on crystal structure analysis, we discuss the regulatory mechanism of MtbAK.  相似文献   

7.
Motoneuron synapses on spinal cord interneurons known as Renshaw cells activate nicotinic, AMPA and NMDA receptors consistent with co-release of acetylcholine and excitatory amino acids (EAA). However, whether these synapses express vesicular glutamate transporters (VGLUTs) capable of accumulating glutamate into synaptic vesicles is controversial. An alternative possibility is that these synapses release other EAAs, like aspartate, not dependent on VGLUTs. To clarify the exact EAA concentrated at motor axon synapses we performed a quantitative postembedding colloidal gold immunoelectron analysis for aspartate and glutamate on motor axon synapses (identified by immunoreactivity to the vesicular acetylcholine transporter; VAChT) contacting calbindin-immunoreactive (-IR) Renshaw cell dendrites. The results show that 71% to 80% of motor axon synaptic boutons on Renshaw cells contained aspartate immunolabeling two standard deviations above average neuropil labeling. Moreover, VAChT-IR synapses on Renshaw cells contained, on average, aspartate immunolabeling at 2.5 to 2.8 times above the average neuropil level. In contrast, glutamate enrichment was lower; 21% to 44% of VAChT-IR synapses showed glutamate-IR two standard deviations above average neuropil labeling and average glutamate immunogold density was 1.7 to 2.0 times the neuropil level. The results were not influenced by antibody affinities because glutamate antibodies detected glutamate-enriched brain homogenates more efficiently than aspartate antibodies detecting aspartate-enriched brain homogenates. Furthermore, synaptic boutons with ultrastructural features of Type I excitatory synapses were always labeled by glutamate antibodies at higher density than motor axon synapses. We conclude that motor axon synapses co-express aspartate and glutamate, but aspartate is concentrated at higher levels than glutamate.  相似文献   

8.
The genes from the thermophilic archaeabacterium Methanococcus jannaschii that code for the putative catalytic and regulatory chains of aspartate transcarbamoylase were expressed at high levels in Escherichia coli. Only the M. jannaschii PyrB (Mj-PyrB) gene product exhibited catalytic activity. A purification protocol was devised for the Mj-PyrB and M. jannaschii PyrI (Mj-PyrI) gene products. Molecular weight measurements of the Mj-PyrB and Mj-PyrI gene products revealed that the Mj-PyrB gene product is a trimer and the Mj-PyrI gene product is a dimer. Preliminary characterization of the aspartate transcarbamoylase from M. jannaschii cell-free extract revealed that the enzyme has a similar molecular weight to that of the E. coli holoenzyme. Kinetic analysis of the M. jannaschii aspartate transcarbamoylase from the cell-free extract indicates that the enzyme exhibited limited homotropic cooperativity and little if any regulatory properties. The purified Mj-catalytic trimer exhibited hyperbolic kinetics, with an activation energy similar to that observed for the E. coli catalytic trimer. Homology models of the Mj-PyrB and Mj-PyrI gene products were constructed based on the three-dimensional structures of the homologous E. coli proteins. The residues known to be critical for catalysis, regulation, and formation of the quaternary structure from the well characterized E. coli aspartate transcarbamoylase were compared.  相似文献   

9.
A permeabilization procedure was adapted to allow the in situ determination of aspartate transcarbamylase activity in Saccharomyces cerevisiae. Permeabilization is obtained by treating cell suspensions with small amounts of 10% toluene in absolute ethanol. After washing, the cells can be used directly in the enzyme assays. Kinetic studies of aspartate transcarbamylase (EC 2.1.3.2) in such permeabilized cells showed that apparent Km for substrates and Ki for the feedback inhibitor UTP were only slightly different from those reported using partially purified enzyme. The aspartate saturation curve is hyperbolic both in the presence and absence of UTP. The inhibition by this nucleotide is noncompetitive with respect to aspartate, decreasing both the affinity for this substrate and the maximal velocity of the reaction. The saturation curves for both substrates give parallel double reciprocal plots. The inhibition by the products is linear noncompetitive. Succinate, an aspartate analog, provokes competitive and uncompetitive inhibitions toward aspartate and carbamyl phosphate, respectively. The inhibition by phosphonacetate, a carbamyl phosphate analog, is uncompetitive and noncompetitive toward carbamyl phosphate and aspartate, respectively, but pyrophosphate inhibition is competitive toward carbamyl phosphate and noncompetitive toward aspartate. These results, as well as the effect of the transition state analog N-phosphonacetyl-L-aspartate, all exclude a random mechanism for aspartate transcarbamylase. Most of the data suggest an ordered mechanism except the substrates saturation curves, which are indicative of a ping-pong mechanism. Such a discrepancy might be related to some channeling of carbamyl phosphate between carbamyl phosphate synthetase and aspartate transcarbamylase catalytic sites.  相似文献   

10.
Sulphydryl groups of mitochondrial aspartate aminotransferase from horse heart were titrated with 5,5'-dithiobis (2-nitrobenzoic acid). From analysis of peptic peptides, 378 amino acid residues (94.3% of the total) in the protein were identified. The results of amino acid sequence analysis are compared with those of cytosolic and mitochondrial aspartate aminotransferases from other sources.  相似文献   

11.
A strain of Escherichia coli has been constructed which greatly overproduces the enzyme aspartate transcarbamylase. This strain has a deletion in the pyrB region of the chromosome and also carries a leaky mutation in pyrF. Although this strain is a pyrimidine auxotroph, it will grow very slowly without pyrimidines if a plasmid containing the pyrB gene is introduced into it. Derepression occurs when this strain exhausts its uracil supply during exponential growth. Under extreme derepression, aspartate transcarbamylase can account for as much as 60% of the total cellular protein. This host strain/plasmid system can be utilized for the rapid purification of wild-type aspartate transcarbamylase or plasmid-born mutant versions of the enzyme. This system is particularly well-suited for analysis of the latter since the control of overproduction resides exclusively on the bacterial chromosome. Therefore, any plasmid bearing the pyrBI operon can be made to overproduce aspartate transcarbamylase in this host strain. Based on this system, a rapid purification procedure has been developed for E. coli aspartate transcarbamylase. The purification scheme involves an ammonium sulfate fractionation followed by a single precipitation of the enzyme at its isoelectric point. In a similar fashion, this strain can also be employed to produce exclusively the catalytic subunit of the enzyme if the plasmid only carries the pyrB gene. This system may be adapted to overproduce other proteins as well by using this host strain and the strong pyrB promoter linked to another gene.  相似文献   

12.
The response regulator AlgR is required for both alginate biosynthesis and type IV fimbria-mediated twitching motility in Pseudomonas aeruginosa. In this study, the roles of AlgR signal transduction and phosphorylation in twitching motility and biofilm formation were examined. The predicted phosphorylation site of AlgR (aspartate 54) and a second aspartate (aspartate 85) in the receiver domain of AlgR were mutated to asparagine, and mutant algR alleles were introduced into the chromosome of P. aeruginosa strains PAK and PAO1. Assays of these mutants demonstrated that aspartate 54 but not aspartate 85 of AlgR is required for twitching motility and biofilm initiation. However, strains expressing AlgR D85N were found to be hyperfimbriate, indicating that both aspartate 54 and aspartate 85 are involved in fimbrial biogenesis and function. algD mutants were observed to have wild-type twitching motility, indicating that AlgR control of twitching motility is not mediated via its role in the control of alginate biosynthesis. In vitro phosphorylation assays showed that AlgR D54N is not phosphorylated by the enteric histidine kinase CheA. These findings indicate that phosphorylation of AlgR most likely occurs at aspartate 54 and that aspartate 54 and aspartate 85 of AlgR are required for the control of the molecular events governing fimbrial biogenesis, twitching motility, and biofilm formation in P. aeruginosa.  相似文献   

13.
Aspartate aminotransferase isoenzymes are located in both the cytosol and organelles of eukaryotes, but all are encoded in the nuclear genome. In the work described here, a phylogenetic analysis was made of aspartate aminotransferases from plants, animals, yeast, and a number of bacteria. This analysis suggested that five distinct branches are present in the aspartate aminotransferase tree. Mitochondrial forms of the enzyme form one distinct group, bacterial aspartate aminotransferase formed another, and the plant and vertebrate cytosolic isoenzymes each formed a distinct group. Plant cytosolic isozymes formed a further group of which the plastid sequences were a member. The yeast mitochondrial and cytosolic aspartate aminotransferases formed groups separate from other members of the family. Correspondence to: C.J. Marshall  相似文献   

14.
A mutant of Rhizobium meliloti, 4R3, which is unable to grow on aspartate has been isolated. The defect is specific to aspartate utilization, since 4R3 is not an auxotroph and grows as well as its parent strain on other carbon and nitrogen sources. The defect was correlated with an inability to fix nitrogen within nodules formed on alfalfa. Transport of aspartate into the mutant cells was found to be normal. Analysis of enzymes involved in aspartate catabolism showed a significantly lower level of aspartate aminotransferase activity in cell extracts of 4R3 than in the wild type. Two unrelated regions identified from a genomic cosmid bank each complemented the aspartate catabolism and symbiotic defects in 4R3. One of the cosmids was found to encode an aspartate aminotransferase enzyme and resulted in restoration of aspartate aminotransferase activity in the mutant. Analysis of the region cloned in this cosmid by transposon mutagenesis showed that mutations within this region generate the original mutant phenotypes. The second type of cosmid was found to encode an aromatic aminotransferase enzyme and resulted in highly elevated levels of aromatic aminotransferase activity. This enzyme apparently compensated for the mutation by its ability to partially utilize aspartate as a substrate. These findings demonstrate that R. meliloti contains an aspartate aminotransferase activity required for symbiotic nitrogen fixation and implicate aspartate as an essential substrate for bacteria in the nodule.  相似文献   

15.
Rapid Inactivation of Brain Glutamate Decarboxylase by Aspartate   总被引:2,自引:2,他引:0  
In the absence of its cofactor, pyridoxal 5'-phosphate (pyridoxal-P), glutamate decarboxylase is rapidly inactivated by aspartate. Inactivation is a first-order process and the apparent rate constant is a simple saturation function of the concentration of aspartate. For the beta-form of the enzyme, the concentration of aspartate giving the half-maximal rate of inactivation is 6.1 +/- 1.3 mM and the maximal apparent rate constant is 1.02 +/- 0.09 min-1, which corresponds to a half-time of inactivation of 41 s. The rate of inactivation by aspartate is about 25 times faster than inactivation by glutamate or gamma-aminobutyric acid (GABA). Inactivation is accompanied by a rapid conversion of holoenzyme to apoenzyme and is opposed by pyridoxal-P, suggesting that inactivation results from an alternative transamination of aspartate catalyzed by the enzyme, as previously observed with glutamate and GABA. Consistent with this mechanism pyridoxamine 5'-phosphate, an expected transamination product, was formed when the enzyme was incubated with aspartate and pyridoxal-P. The rate of transamination relative to the rate of decarboxylation was much greater for aspartate than for glutamate. Apoenzyme formed by transamination of aspartate was reactivated with pyridoxal-P. In view of the high rate of inactivation, aspartate may affect the level of apoenzyme in brain.  相似文献   

16.
Each of two previously isolated strains of Escherichia coli containing a single nonsense codon within the pyrB gene was suppressed with four different nonsense suppressors. The kinetic analysis using crude extracts of these nonsense-suppressed strains indicated that the mutant aspartate transcarbamylases had altered cooperativity and affinity for aspartate as judged by the substrate concentration at half of the maximal velocity. Both pyrB genes were cloned and then sequenced. In both cases, a single base change was identified which converted a glutamine GAC codon into a TAC nonsense codon. Both mutations occurred in the catalytic chain of aspartate transcarbamylase and were identified at positions 108 and 246. The glutamine at position 108 in the wild-type structure is located at the interface between the catalytic and regulatory chains and is involved in a number of interactions with backbone and side chains of the regulatory chain. The glutamine at position 246 in the wild-type structure is located in the 240s loop of the enzyme. Two additional mutant versions of aspartate transcarbamylase were created by site-directed mutagenesis to further investigate the 108-position in the structure, a glutamine to tyrosine substitution at position 108 of the catalytic chain, and an asparagine to glycine change at position 113 of the regulatory chain, a residue which interacts directly with glutamine-108 in the wild-type structure. Both mutant enzymes have reduced affinity for aspartate. However, the Tyr-108 mutant enzyme exhibits a reduced Hill coefficient while the Gly-113 enzyme exhibits an increased Hill coefficient. The response to the allosteric effectors ATP and CTP is also changed for both the mutant enzymes.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
In brain the amino acid L-aspartate serves roles as: (1) putative transmitter, (2) protein precursor, (3) donor of atoms for the biosynthesis of pyrimidine and purine bases, and (4) fuel for energy metabolism. Astrocytes dominate aspartate clearance in brain, and in culture they take up aspartate and quickly metabolize it. In brain, only astrocytes were shown to express the enzymes for de novo pyrimidine biosynthesis. To gain more details about the spectrum of metabolites generated from aspartate and subsequently released by cultured astrocytes a 13C-nuclear magnetic resonance analysis was performed of [U-13C]aspartate supplemented incubation media exposed to astroglial cultures. The results show that astrocytes readily metabolize aspartate and release into their culture media 13C-isotopomers of lactate, glutamine, citrate and alanine. Despite the presence in astroglial cells of two tandem enzymes of pyrimidine biosynthesis and their mRNAs, pyrimidine nucleotide-related heterocyclic compounds such as dihydroorotate and orotate could not be detected in the culture media.  相似文献   

18.
The use of [14C]aspartate of high specific activity and thin-layer chromatography on polyethyleneimine cellulose for the separation of carbamoyl aspartate from aspartate has enabled the measurement of aspartate carbamoyltransferase and carbamoyl phosphate synthase activities and carbamoyl phosphate concentrations in extracts from Escherichia coli. The assay method described is sensitive to the formation of about 1 pmol of carbamoyl aspartate.  相似文献   

19.
A method has been developed which allows isolation of 0.3--0.5 g of mitochondrial aspartate aminotransferase in five days starting from 10 pig hearts; the method does not involve initial preparation of mitochondria. Mitochondrial malate dehydrogenase and the cytoplasmic aspartate aminotransferase may conveniently be recovered from side fractions. The product mitochondrial aspartate aminotransferase is homogeneous as judged by various electrophoretic techniques and by N-terminal analysis. Crystals of the enzyme have been obtained both from concentrated, essentially salt-free, solutions and from solutions of ammonium sulphate. The amino acid composition, N and C-terminal amino acid sequences and subunit molecular weight have been determined; these characteristic properties are compared with those of the cytoplasmic isozyme from the same source.  相似文献   

20.
Multiplicity of aspartate transport in thin wastewater biofilms.   总被引:3,自引:3,他引:0       下载免费PDF全文
This research documents the multiplicity of L-aspartate transport in thin wastewater biofilms. A Line-weaver-Burk analysis of incorporation produced a curvilinear plot (concave down) that suggested active transport by two distinct systems (1 and 2). The inactivation of system 2 with AsO4 or osmotic shock resolved system 1, which was a high-affinity, low-capacity system with an apparent Kt (Michaelis-Menten constant) of 4.3 microM (AsO4) or 4.6 microM (osmotic shock). The inactivation of system 1 with dinitrophenol resolved system 2, which was a low-affinity, high-capacity system with an apparent Kt of 116.7 microM. System 1 was more specific for aspartate than system 2 in the presence of aspartate analogs. Sodium had no discernible effect on the incorporation velocities by either system. These results indicate that system 1 is a membrane-bound proton symport coupled to the proton gradient component of the proton motive force and that system 2 is a binding protein-mediated system coupled to phosphate bond energy. Analyses of diffusional limitations on the derived transport constants indicated that internal resistances were present but that the apparent constants were close to the intrinsic values, especially for system 1. Metabolic inactivation of the biofilm with dinitrophenol and AsO4 did not completely inactivate aspartate incorporation, which indicated that some simple adsorption of the aspartate anion by the biofilm had occurred. These results show that aspartate is transported by wastewater biofilm bacteria via systems with different affinities, specificities, and mechanisms of energy coupling.  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