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 78 毫秒
1.
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   

2.
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   

3.
The urea cycle in the liver of adjuvant-induced arthritic rats was investigated using the isolated perfused liver. Urea production in livers from arthritic rats was decreased during substrate-free perfusion and also in the presence of the following substrates: alanine, alanine + ornithine, ammonia, ammonia + lactate, ammonia + pyruvate and glutamine but increased when arginine and citrulline + aspartate were the substrates. No differences were found with ammonia + aspartate, ammonia + aspartate + glutamate, aspartate, aspartate + glutamate and citrulline. Ammonia consumption was smaller in the arthritic condition when the substance was infused together with lactate or pyruvate but higher when the substance was simultaneously infused with aspartate or aspartate + glutamate. Glucose production tended to correlate with the smaller or higher rates of urea synthesis. Blood urea was higher in arthritic rats (+25.6%), but blood ammonia was lower (–32.2%). Critical for the synthesis of urea from various substrates in arthritic rats seems to be the availability of aspartate, whose production in the liver is probably limited by both the reduced gluconeogenesis and aminotransferase activities. This is indicated by urea synthesis which was never inferior in the arthritic condition when aspartate was exogenously supplied, being even higher when both aspartate and citrulline were simultaneously present. Possibly, the liver of arthritic rats has a different substrate supply of nitrogenous compounds. This could be in the form of different concentrations of aspartate or other aminoacids such as citrulline or arginine (from the kidneys) which allow higher rates of hepatic ureogenesis.  相似文献   

4.
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   

5.
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   

6.
A major pathway of beta-alanine synthesis in insects is through the alpha-decarboxylation of aspartate, but the enzyme involved in the decarboxylation of aspartate has not been clearly defined in mosquitoes and characterized in any insect species. In this study, we expressed two putative mosquito glutamate decarboxylase-like enzymes of mosquitoes and critically analyzed their substrate specificity and biochemical properties. Our results provide clear biochemical evidence establishing that one of them is an aspartate decarboxylase and the other is a glutamate decarboxylase. The mosquito aspartate decarboxylase functions exclusively on the production of beta-alanine with no activity with glutamate. Likewise the mosquito glutamate decarboxylase is highly specific to glutamate with essentially no activity with aspartate. Although insect aspartate decarboxylase shares high sequence identity with glutamate decarboxylase, we are able to closely predict aspartate decarboxylase from glutamate decarboxylase based on the difference of their active site residues.  相似文献   

7.
13C kinetic isotope effects have been measured in carbamyl phosphate for the reaction catalyzed by aspartate transcarbamylase. For the holoenzyme, the value was 1.0217 at zero aspartate, but unity at infinite aspartate, with 4.8 mM aspartate eliminating half of the isotope effect. This pattern proves an ordered kinetic mechanism, with carbamyl phosphate adding before aspartate. The same parameters were observed in the presence of ATP or CTP, showing that there is only one form of active enzyme present, regardless of the presence or absence of allosteric modifiers. These data support the Monod model of allosteric behavior in which the equilibrium between fully active and inactive enzyme is perturbed by selective binding interactions of substrates and modifiers, and there are no enzyme forms having partial activity. Isolated catalytic subunits of the enzyme showed similar 13C isotope effects (1.0240 at zero aspartate, 1.0039 at infinite aspartate, 3.8 mM aspartate causing half of the change from one value to the other), but the finite isotope effect at infinite aspartate shows that the kinetic mechanism is now partly random. With the very slow and poorly bound aspartate analog cysteine sulfinate, the 13C isotope effects were 1.039 for both holoenzyme and catalytic subunits and were not decreased significantly by high levels of cysteine sulfinate. The value of 1.039 is probably close to the intrinsic isotope effect on the chemical reaction, while the kinetic mechanism with this substrate is now fully random because the chemistry is so much slower than release of either reactant from the enzyme.  相似文献   

8.
A high resolution structure of the N-terminal ligand-binding domain of the aspartate receptor which mediates aspartate chemotaxis in Salmonella typhimurium has recently been reported. A least-squares superposition of the alpha-amino nitrogen, alpha-carbon, beta-carbon, and alpha-carboxylate carbon of the aspartate bound to the aspartate receptor onto the equivalent atoms in the tryptophan bound to the trp repressor provides evidence for similarity between key parts of the active sites that bind to the alpha-amino and alpha-carboxylates of the respective ligands. Because the N-terminal domain of the aspartate receptor and the trp repressor also share other structural similarities, we hypothesize that the similarity between the aspartate receptor and the trp repressor derives from a similarity in ligand-induced conformational changes at the active sites of these proteins. This hypothesis also implies that an important signaling event in the aspartate receptor occurs through tertiary conformational changes within a single subunit.  相似文献   

9.
The periplasmic domain of the aspartate chemotaxis receptor from Salmonella typhimurium has been crystallized in the presence and absence of bound aspartate. Both crystal forms were grown by precipitation with lithium sulfate and diffract to 1.8 A resolution. The aspartate receptor structure is believed to be prototypical of a large class of receptors including those for polypeptide growth factor hormones as well as those for small chemotaxis-affector molecules such as aspartate and serine.  相似文献   

10.
Activity of aspartate aminotransferase, an enzyme which catalyzes the interconversion of the excitatory transmitter candidates, glutamate and aspartate, has been measured in fiber tracts of rat, with an emphasis on sensory and motor systems of the brain. Most tracts had significantly higher activities than the cholinergic facial nerve root, consistent with the possibility that a component of aspartate aminotransferase activity might serve as a marker for neurons using glutamate and/or aspartate as neurotransmitter. Highest activity was in the auditory nerve root. On the other hand, a close correlation was found between aspartate aminotransferase and malate dehydrogenase activities in the fiber tracts, raising the question whether aspartate aminotransferase activity may be more closely related to energy metabolism than to transmitter metabolism.  相似文献   

11.
In accordance with the regulation by aspartate of phosphoenolpyrubate (PEP*) carboxylase, glutamate formation in Brevibacterium flavum, a glutamate-producing bacterium, was inhibited by the addition of aspartate. Furthermore, an increase in aspartate formation caused by a mutational decrease in citrate synthase specific activity was accompanied by a decrease in the total amount of glutamate and aspartate formed. However, a mutational decrease in glutamate dehydrogenase activity caused a decrease in the total amount without increasing the asparate formation but with accumulation of 2-oxoglutarate, suggesting that the feedback inhibition by the aspartate of PEP carboxylase was enhanced by 2-oxoglutarate. In fact, partially purified PEP carboxylase from this organism was found to be synergistically inhibited by aspartate and 2-oxoglutarate, citrate, cis-aconitase, or isocitrate. Among them, the effects of tricarboxylic acids were attributed to their non-specific chelating action with Mn2+, an activator of the enzyme. The synergistic action of 2-oxoglutarate was accompanied by a decrease in Hill coefficient for the aspartate of the enzyme.  相似文献   

12.
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   

13.
A search in the Bacillus subtilis genome sequence found that the gene designated yclM encode(s) a protein showing significant identity in amino acid sequence to aspartate kinases. When yclM was introduced into Escherichia coli cells deficient in all three aspartate kinase genes, production of a protein with molecular size 50 kDa, which was similar to the value deduced from the nucleotide sequence of the gene, was observed. Expectedly, the protein purified to homogeneity had aspartate kinase activity. The enzyme was significantly inhibited by simultaneous addition of both threonine and lysine, which is a typical feature of aspartate kinase III of B. subtilis. The enzyme was very unstable in 10 mM tris-HCl (pH 7.5) buffer, but was stabilized by addition of 500 mM ammonium sulfate. Although all the aspartate kinases so far investigated are oligomeric enzymes, this aspartate kinase was suggested to be a monomer.  相似文献   

14.
The high-resolution structures of the wild-type periplasmic domain of the bacterial aspartate receptor have been determined in the absence and presence of bound aspartate to 1.85 and 2.2 Å resolution, respectively. As we reported earlier, in the refined structure of the complexed form of the crosslinked cysteine mutant receptor, the binding of the aspartate at the first site was mediated through four bridging water molecules while the second site showed an occupant electron density that best fit a sulfate group, which was present in the crystallization solution at high concentration. In the wild-type periplasmic domain structure two aspartate residues are bound per dimer, but with different occupancies. There exists a “strong” aspartate-binding site whose binding is again mediated by four water molecules while the second site contains aspartate whoseB-factor is about 10% higher, signifying weaker binding. The interaction between the second, “weaker” aspartate with the three ligand-binding arginine side-chains is slightly different from the first site. The major difference is that there are three water molecules mediating the binding of aspartate at the second site, whereas in the first site there are four bridging water molecules. The fact that aspartate-complexed crystals of the wild-type were grown with a large excess aspartate while the cross-linked crystals were grown with equal molar aspartate may explain the difference in the stoichiometry observed. The conservation of the four bridging water molecules in the strong aspartate site of both the cross-linked and wild-type periplasmic domain may reflect an important binding motif.The periplasmic domain in the apo form is a symmetrical dimer, in which each of the subunits is equivalent, and the two aspartate binding sites are identical. Upon the binding of aspartate, the subunits are no longer symmetrical. The main difference between the aspartate-bound and unbound forms is in a small, rigid-body rotation between the subunits within a dimer. The rotation is similar in both direction and magnitude in the crosslinked and wild-type periplasmic domains. The presence of the second aspartate in the wild-type structure does not make any additional rotation compared to the single-site binding. The conservation of the small angular changein vitrosuggests that the inter-subunit rotation may have relevance to the understanding of the mechanism of transmembrane signal transductionin vivo.  相似文献   

15.
The aspartate/glutamate carrier from bovine heart mitochondria was solubilized with dodecyl-octaoxyethylene ether (C12E8) and purified by chromatography on hydroxyapatite and celite. On SDS gel electrophoresis, the purified aspartate/glutamate carrier consisted of a single protein band with an apparent Mr of 31,500. When reconstituted into liposomes the aspartate/glutamate carrier protein catalyzed an N-ethylmaleimide-sensitive aspartate/aspartate exchange. It was purified 620-fold with a recovery of 17.2% and a protein yield of 0.03% with respect to the mitochondrial extract. The properties of the reconstituted carrier, i.e. requirement for a counteranion, substrate specificity and inhibitor sensitivity, were similar to those of the aspartate/glutamate carrier as characterized in mitochondria.  相似文献   

16.
Aspartate stimulated by as much as three fold the rate of malate decarboxylation by Zea mays bundle sheath cells. Both the basal and aspartate stimulated rates of malate decarboxylation were light-dependent. Stimulation appeared to be due to aspartate as such, rather than depending on aspartate metabolism, and was due partly to a reduction in the malate concentration required for maximum decarboxylation and partly to an increased maximum velocity of decarboxylation. The extractable activities of NADP malic enzyme, glyceraldehyde phosphate dehydrogenase, and 3-phosphoglycerate kinase recoverable from cells were not increased by preincubating cells with aspartate, and aspartate did not affect the activity of these enzymes in cell-free extracts. It is suggested that aspartate may influence the transport of either malate into or pyruvate out of bundle sheath chloroplasts.  相似文献   

17.
In previous studies it was found that: (a) aspartate aminotransferase increases the aspartate dehydrogenase activity of glutamate dehydrogenase; (b) the pyridoxamine-P form of this aminotransferase can form an enzyme-enzyme complex with glutamate dehydrogenase; and (c) the pyridoxamine-P form can be dehydrogenated to the pyridoxal-P form by glutamate dehydrogenase. It was therefore concluded (Fahien, L.A., and Smith, S.E. (1974) J. Biol. Chem 249, 2696-2703) that in the aspartate dehydrogenase reaction, aspartate converts the aminotransferase into the pyridoxamine-P form which is then dehydrogenated by glutamate dehydrogenase. The present results support this mechanism and essentially exclude the possibility that aspartate actually reacts with glutamate dehydrogenase and the aminotransferase is an allosteric activator. Indeed, it was found that aspartate is actually an activator of the reaction between glutamate dehydrogenase and the pyridoxamine-P form of the aminotransferase. Aspartate also markedly activated the alanine dehydrogenase reaction catalyzed by glutamate dehydrogenase plus alanine aminotransferase and the ornithine dehydrogenase reaction catalyzed by ornithine aminotransferase plus glutamate dehydrogenase. In these latter two reactions, there is no significant conversion of aspartate to oxalecetate and other compounds tested (including oxalacetate) would not substitute for aspartate. Thus aspartate is apparently bound to glutamate dehydrogenase and this increases the reactivity of this enzyme with the pyridoxamine-P form of aminotransferases. This could be of physiological importance because aspartate enables the aspartate and ornithine dehydrogenase reactions to be catalyzed almost as rapidly by complexes between glutamate dehydrogenase and the appropriate mitochondrial aminotransferase in the absence of alpha-ketoglutarate as they are in the presence of this substrate. Furthermore, in the presence of aspartate, alpha-ketoglutarate can have little or no affect on these reactions. Consequently, in the mitochondria of some organs these reactions could be catalyzed exclusively by enzyme-enzyme complexes even in the presence of alpha-ketoglutarate. Rat liver glutamate dehydrogenase is essentially as active as thebovine liver enzyme with aminotransferases. Since the rat liver enzyme does not polymerize, this unambiguously demonstrates that monomeric forms of glutamate dehydrogenase can react with aminotransferases.  相似文献   

18.
Tyrosine, added to the growth medium of a strain of Escherichia coli K-12 lacking transaminase B, repressed the tyrosine, phenylalanine, and tryptophan aminotransferase activities while leaving the aspartate aminotransferase activity unchanged. This suggested that the aspartate and the aromatic aminotransferase activities, previously believed to reside in the same protein, viz. transaminase A, are actually nonidentical. Further experiments showed that, upon incubation at 55 C, the aspartate aminotransferase of crude extracts was almost completely stable, whereas the tyrosine and phenylalanine activities were rapidly inactivated. Apoenzyme formation was faster, and apoenzyme degradation proceeded more slowly with aspartate aminotransferase than with tyrosine aminotransferase. Electrophoresis in polyacrylamide gels separated the aminotransferases. A more rapidly moving band contained tyrosine, phenylalanine, and tryptophan aminotransferases, and a slower band contained aspartate aminotransferase. A mutant of E. coli K-12 with low levels of aspartate aminotransferase exhibited unchanged levels of tyrosine aminotransferase. Thus, transaminase A appears to be made up of at least two proteins: one of broad specificity whose synthesis is repressed by tyrosine and another, specific for aspartate, which is not subject to repression by amino acids. The apparent molecular weights of both the aspartate and the aromatic aminotransferases, determined by gel filtration, were about 100,000.  相似文献   

19.
In order to study the mechanism of the glutamate-aspartate translocator, rat liver mitochondria were loaded with either glutamate or aspartate. In the presence of ascorbate plus tetramethyl-p-phenylenediamine as an electron donor at the third energy conservation site, exchange of external glutamate for matrix aspartate is highly favored over the reverse exchange. In the absence of an energy source, although the asymmetry of the exchange rates is much smaller, it is still observable. Further studies have shown that the proton uptake accompanying influx of glutamate in exchange for aspartate efflux occurs by protonation of a group on the carrier (pK = 7.9) at the external side of the inner mitochondrial membrane, followed by deprotonation at the matrix surface. It is postulated that glutamate binds to the protonated form of the carrier and aspartate to the deprotonated form. Because of the alkaline pK, aspartate efflux is inhibited with increased matrix [H+] due to limitation of the availability of deprotonated carrier for aspartate binding. For the reverse exchange, aspartate uptake is inhibited by increasing external [H+]. Thus the rate of aspartate uptake by mitochondria is apparently impeded both by a proton motive force (Δp) unfavorable to entry of ions with net negative charge as well as by the small proportion of deprotonated carrier at the outer surface of the membrane. This conclusion is further illustrated by inhibition of the aspartate-aspartate exchange with increased [H+] and by addition of an energy source. The glutamate-glutamate exchange, however, showed a slight stimulation by increased [H+] and was unaffected by the energy state.The model initially proposed for the carrier, in which a neutral glutamate-carrier complex exchanges for a negatively charged aspartate-carrier complex, is tested further. Glutamate uptake was noncompetitively inhibited by external aspartate, which indicates that aspartate and glutamate bind to separate forms of the carrier. Intramitochrondrial glutamate at a concentration of 18 mm, however, had no effect on aspartate efflux. Arrhenius plots for the glutamate-aspartate and aspartate-glutamate exchanges were linear over the range of temperatures tested (1–35 °C and 5–25 °C, respectively) and provided an average value of 14.3 kcal/mol for the energy of activation. In addition, there appear to be two pools, exchangeable and nonexchangeable, of matrix aspartate available to the translocator, since extramitochondrial radiolabeled aspartate can equilibrate only with unlabeled matrix aspartate at low aspartate loading (1–2 nmol of aspartate/mg of protein). The physiological significance of the data is discussed.  相似文献   

20.
PS1 deficiency and expression of PS1 with substitutions of two conserved transmembrane aspartate residues ("PS1 aspartate variants") leads to the reduction of Abeta peptide secretion and the accumulation of amyloid precursor protein (APP) C-terminal fragments. To define the nature of the "dominant negative" effect of the PS1 aspartate variants, we stably expressed PS1 harboring aspartate to alanine substitutions at codons 257 (D257A) or 385 (D385A), singly or in combination (D257A/D385A), in mouse neuroblastoma, N2a cells. Expression of the PS1 aspartate variants resulted in marked accumulation of intracellular and cell surface APP C-terminal fragments. While expression of the D385A PS1 variant reduced the levels of secreted Abeta peptides, we now show that neither the PS1 D257A nor D257A/D385A variants impair Abeta production. Surprisingly, the stability of both immature and mature forms of APP is dramatically elevated in cells expressing PS1 aspartate variants, commensurate with an increase in the cell surface levels of APP. These findings lead us to conclude that the stability and trafficking of APP can be profoundly modulated by coexpression of PS1 with mutations at aspartate 257 and aspartate 385.  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