首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The 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   

2.
A simple and rapid affinity chromatographic method for the isolation of aspartate transcarbamylase from germinated seedlings of mung bean (Phaseolus aureus) was developed. A partially purified preparation of the enzyme was chromatographed on an affinity column containing aspartate linked to CNBr-activated Sepharose 4B. Aspartate transcarbamylase was specifically eluted from the column with 10 mm aspartate or 0.5 m KCl. The enzyme migrated as a single sharp band during disc electrophoresis at pH 8.6 on polyacrylamide gels. Electrophoresis of the sodium dodecyl sulfate-treated enzyme showed two distinct protein bands, suggesting that the mung bean aspartate transcarbamylase was made up of nonidentical subunits. Like the enzyme purified by conventional procedures, this enzyme preparation also exhibited positive homotropic interactions with carbamyl phosphate and negative heterotropic interactions with UMP. This method was extended to the purification of aspartate transcarbamylase from Lathyrus sativus, Eleucine coracona, and Trigonella foenum graecum.  相似文献   

3.
Modes of modifier action in E. coli aspartate transcarbamylase   总被引:4,自引:0,他引:4  
The observed patterns for inhibition by CTP and succinate of equilibrium exchange kinetics with native aspartate transcarbamylase (E. coli) are consistent with an ordered substrate-binding system in which aspartate binds after carbamyl phosphate, and phosphate is released after carbamyl aspartate. ATP selectively stimulates Asp carbamyl-Asp exchange, but not carbamyl phosphate Pi. Initial velocity studies at 5 °, 15 °, and 35 °C were carried out, using modifiers as perturbants of the system. Modifiers alter the Hill n and S0.5 for aspartate, most markedly at 15 °C but less so at the other temperatures. ATP does increase V under saturating substrate conditions, and substrate inhibition is observed for aspartate. ATP does not make the Hill n = 1 at any temperature. It is proposed that CTP and ATP act by separate mechanisms, not by simply perturbing in opposite directions the equilibrium for aspartate binding. ATP appears to act to increase the rate of aspartate association and dissociation, whereas CTP induces an intramolecular competitive effect in the protein.  相似文献   

4.
Aspartate kinase and two homoserine dehydrogenases were partially purified from 4-day-old pea seedlings. A sensitive method for measuring aspartate kinase activity is described. Aspartate kinase activity was dependent upon ATP, Mg2+ or Mn2+, and aspartate. The aspartate kinase was inhibited in a sigmoidal manner by threonine and Ki for threonine was 0·57 mM. The enzyme could be desensitized to the inhibitor and threonine protected the enzyme against thermal inactivation. Aspartate kinase activity was enhanced by isoleucine, valine and alanine. Homoserine, methionine and lysine were without effect. The homoserine dehydrogenase activity which was associated with aspartate kinase during purification could be resolved into two peaks by gel filtration. The activity of both peaks was inhibited by aspartate and cysteine and one was inhibited by threonine.  相似文献   

5.
The mitochondrial glutamate-aspartate exchange carrier catalyzes the electrogenic exchange of intramitochondrial aspartate for extramitochondrial glutamate. Protons are cotransported with glutamate in a 1:1 ratio. In the present study, the effects of pH and glutamate concentration on glutamate entry into intact mitochondria were determined. Hydrogen ions were found to decrease the Km for glutamate entry. In addition, using glutamate-loaded submitochondrial particles, aspartate transport into the particles was measured as a function of internal and external glutamate concentrations, pH, and electrical potential across the membrane. Glutamate, was a competitive inhibitor of aspartate transport when both amino acids were present on the same side of the membrane, while H+ was a noncompetitive inhibitor of aspartate entry into the particles. A decrease in glutamate concentration on the inside of the particles brought about a parallel decrease in V and Km for aspartate outside of the particles, thus suggesting a ping-pong mechanism for the carrier. The uncoupling agent, carbonyl cyanide p-trifluoromethoxy-phenylhydrazone (FCCP), lowered both the Km and V of aspartate transport, while the effect on V was somewhat larger. Data obtained in the presence of KSCN was similar to that obtained with FCCP, and therefore it is concluded that both Km and V changes are dependent on a change of electrical potential across the membrane. A model for the carrier is proposed, which is consistent with the data presented. The model includes a single binding site specific for either glutamate or aspartate, and a separate binding site for the cotransported proton. The affinity of the binding site for protons is increased by simultaneous glutamate binding, but decreased by aspartate binding. The data suggest that an increase in the membrane potential increases the mobility of the charged carrier-aspartate complex, but also facilitates some additional step in the exchange cycle involving subsequent return of the carrier to the matrix side of the membrane. The additional membrane-potential-dependent step could be proton binding on the cytosolic side of the carrier.  相似文献   

6.
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   

7.
Some kinetic studies of the interactions between Escherichia coli phosphoenolpyruvate carboxylase (orthophosphate:oxaloacetate carboxylase (phosphorylating) EC 4.1.1.31) acetyl coenzyme A, fructose 1,6-bisphosphate, and aspartate were performed. Activation of the enzyme by fructose 1,6-bisphosphate is anomalous by comparison with acetyl coenzyme A in that it confers hysteretic properties on the enzyme. In the presence of both activators and aspartate, hysteresis is observed also, but the approach to optimum catalytic activity can be fit to an equation for a second-order reaction with respect to enzyme concentration. Since, however, hysteresis is not a result of any apparent association-dissociation reaction, the apparent fit to a second-order kinetic equation is probably not real but is the result of a multistep activation mechanism. Hysteresis is not eliminated by preincubation of the enzyme with fructose 1,6-bisphosphate, acetyl coenzyme A, or phosphoenolpyruvate singly or in any pair of combinations. Hysteresis is associated, therefore, with the slow conformation change from the inactive species to the active species under the influence of all three of those reactants. The enzyme complex resulting from the binding of each activator, including phosphoenolpyruvate, has an increased affinity for the other activators. A kinetic method for estimating the relative changes in affinity of these complexes for some of the other reactants is presented. At concentrations of the activators below their Ka, synergistic effects are evident, particularly in their ability to relieve aspartate inhibition. Aspartate inhibition is competitive with acetyl coenzyme A both in the absence and in the presence of low concentrations of fructose 1,6-bisphosphate. Increasing the concentrations of fructose 1,6-bisphosphate results in an increase in the apparent Kl for aspartate, suggesting that synergistic activation by fructose 1,6-bisphosphate is a result of the increased affinity of the fructose 1,6-bisphosphate-enzyme complex for acetyl coenzyme A, and a shift in the concentration of enzyme species away from the one(s) to which aspartate can bind most easily. In the presence of fructose 1,6-bisphosphate alone optimal activation can be achieved, but the concentrations required in vitro are high and suggest that fructose 1,6-bisphosphate alone does not function in that capacity physiologically, but primes the enzyme for more effective activation by acetyl coenzyme A and/or phosphoenolpyruvate.  相似文献   

8.
In vivo and in vitro studies on asparagine biosynthesis in soybean seedlings   总被引:15,自引:0,他引:15  
The biosynthesis of asparagine in plants was investigated by feeding radioactive metabolites to soybean cotyledons and by extracting an asparagine synthetase from the same tissue. Soybean cotyledon slices were supplied with radioactive succinate, malate, or aspartate in the presence or absence of various unlabeled metabolites for periods of up to 80 min. Neither aspartate nor malate was rapidly converted to asparagine; labeled aspartate was converted largely to malate. Labeled succinate was rapidly converted to asparagine, and several lines of evidence suggested that fumarate, malate, and aspartate are intermediates. The results suggest that asparagine biosynthesis in plant cells is compartmentalized beginning with succinate. Although results were also consistent with asparagine formation via aspartate, metabolism of a mixture of [14C] plus [3H]succinate resulted in a lower 14C3H ratio in asparagine than aspartate, suggesting that some asparagine may be formed via another pathway. Demonstration of a glutamine-linked asparagine synthetase in soybean cotyledons supports the idea that asparagine is formed via aspartate. The enzyme requires aspartate (Km = 2.2 mm), glutamine (Km = 0.12 mm), ATP (Km = 0.066 mm), magnesium ion, and sulfhydryl protection. It has a pH optimum of 7.7 and is not located in mitochrondria. A small amount of asparagine was formed when ammonium ion was substituted for glutamine, but the Km of the enzyme for ammonium ion was about 25-fold greater than the Km for glutamine suggesting that glutamine is the physiologically important substrate. Soybean cotyledons actively convert [14C]-cyanide to asparagine, apparently via β-cyanoalanine. However, malate was also rapidly labeled from [14C]cyanide and this result cannot be explained by known metabolic pathways.  相似文献   

9.
An improved colorimetric assay for ornithine and aspartate transcarbamylase has been devised. The conventional method of L. M. Prescott and M. E. Jones (1969, Anal. Biochem.32, 408–419) for the detection of ureido compounds, has been optimized and standardized to a highly reproducible, sensitive, efficient, and inexpensive method for the assay of carbamyl aspartate or citrulline, the products of aspartate transcarbamylase and ornithine transcarbamylase, respectively.  相似文献   

10.
11.
A procedure for determining the activity of aspartate transcarbamylase, based upon the greater ultraviolet absorbancy of the products of the reaction catalyzed compared to the reactants, was devised. Extinction coefficients were determined at 205, 210, and 215 nm for the compounds carbamoyl aspartate, acetyl aspartate, and aspartate. These values formed the quantitative basis for a spectrophotometric assay in which an enzymatic reaction is monitored at one of these wavelengths. Use of this procedure was illustrated in four kinetic experiments with the allosteric aspartate transcarbamylase from Escherichia coli, and the nonallosteric catalytic subunit of this enzyme: aspartate saturation curve, arsenate saturation curve (reverse reaction), allosteric activation by a transition-state analog employing acetyl phosphate as substrate, and carbamoyl phosphate progress curve (substrate depletion in the presence of excess cosubstrate). Owing to changes in absorbance on the order of 1000 liter mol?1 cm?1 concomitant with the reaction, the sensitivity of the method is comparable to that of many procedures already in the literature.  相似文献   

12.
The reaction mechanism of aspartate transcarbamylase from mouse spleen has been determined, using steady-state kinetics, isotope-exchange experiments, inhibition studies with a transition-state analog, and product-inhibition studies. Intersecting reciprocal plots obtained when one substrate was varied against different concentrations of the second substrate indicate that the mechanism is sequential. The transition-state analog, N-(phosphonacetyl)-l-aspartate, was a powerful inhibitor of aspartate transcarbamylase, with an inhibition constant (Ki) of 2.6 × 10?8m at 37 °C and pH 7.4 in 0.05 m Na HEPES buffer. PALA gave competitive inhibition with carbamyl phosphate and noncompetitive inhibition with l-aspartate, indicating that carbamyl phosphate must bind before aspartate for catalysis to occur. A ping-pong mechanism in which carbamyl phosphate binds first was excluded by isotope-exchange experiments, since [32P]inorganic phosphate was not incorporated into carbamyl phosphate in the absence of aspartate. Product-inhibition studies showed that only inorganic phosphate and carbamyl phosphate gave a competitive pattern; all other combinations of substrate and product gave noncompetitive inhibition patterns when incubations were carried out at subsaturating concentrations of the second substrate. These inhibition patterns showed that carbamyl phosphate binds first, aspartate binds second, carbamyl aspartate dissociates first, and phosphate dissociates second.  相似文献   

13.
Arabidopsis possesses several genes encoding aspartate aminotransferase, which catalyzes the bidirectional conversion of aspartate into glutamate. These amino acids together with asparagine and glutamine play an important role in N storage and distribution. In addition, they act as precursors for other amino acids. The gene encoding cytosolic aspartate aminotransferase, Asp2, was found to be induced upon infection with the necrotrophic pathogen Botrytis cinerea in Arabidopsis. Asp2 over-expression lines and a T-DNA insertion mutant were used to study the role of aspartate aminotransferase in Arabidopsis defence responses. Over-expression of Asp2 led to changes in aspartate content and aspartate-derived amino acids. The Asp2 knockout mutant was also slightly affected in its amino acid composition. Under standard growth conditions, the Asp2 transgenic lines did not show morphological changes in comparison with the wild-type. However, transgenic lines with the highest Asp2 expression displayed more spreading lesions when infected with B. cinerea. We discuss how this gene involved in amino acid metabolism might interact with plant defence responses.  相似文献   

14.
More than 90% of the aspartate in a defined medium was metabolized after lactate exhaustion such that 3 mol of aspartate and 1 mol of propionate were converted to 3 mol of succinate, 3 mol of ammonia, 1 mol of acetate, and 1 mol of CO2. This pathway was also evident when propionate and aspartate were the substrates in complex medium in the absence of lactate. In complex medium with lactate present, about 70% of the aspartate was metabolized to succinate and ammonia during lactate fermentation, and as a consequence of aspartate metabolism, more lactate was fermented to acetate and CO2 than was fermented to propionate. The conversion of aspartate to fumarate and ammonia by the enzyme aspartase and subsequent reduction of fumarate to succinate occurred in the five strains of Propionibacterium freudenreichii subsp. shermanii studied. The ability to metabolize aspartate in the presence of lactate appeared to be related to aspartase activity. The specific activity of aspartase increased during and after lactate utilization, and the levels of this enzyme were lower in cells grown in defined medium than levels in those cells grown in complex medium. Under the conditions used, no other amino acids were readily metabolized in the presence of lactate. The possibility that aspartate metabolism by propionibacteria in Swiss cheese has an influence on CO2 production is discussed.  相似文献   

15.
Functional and structural characterizations of pyridoxal 5′-phosphate-independent aspartate racemase of the acidothermophilic archaeon Picrophilus torridus were performed. Picrophilus aspartate racemase exhibited high substrate specificity to aspartic acid. The optimal reaction temperature was 60 °C, which is almost the same as the optimal growth temperature. Reflecting the low pH in the cytosol, the optimal reaction pH of Picrophilus aspartate racemase was approximately 5.5. However, the activity at the putative cytosolic pH of 4.6 was approximately 6 times lower than that at the optimal pH of 5.5. The crystal structure of Picrophilus aspartate racemase was almost the same as that of other pyridoxal 5′-phosphate -independent aspartate racemases. In two molecules of the dimer, one molecule contained a tartaric acid molecule in the catalytic site; the structure of the other molecule was relatively flexible. Finally, we examined the intracellular existence of d-amino acids. Unexpectedly, the proportion of d-aspartate to total aspartate was not very high. In contrast, both d-proline and d-alanine were observed. Because Picrophilus aspartate racemase is highly specific to aspartate, other amino acid racemases might exist in Picrophilus torridus.  相似文献   

16.
Twenty-one monoclonal antibodies were raised against the aspartate aminotransferase-P2 isoenzyme from root nodules of Lupinus angustifolius [L.] cv Uniharvest. Induction of this isoenzyme is positively correlated with the onset of N2 fixation in effective root nodules and is associated with the assimilation of ammonia by the plant in the Rhizobium-legume symbiosis. The monoclonal antibodies produced were all of the IgG class, recognized five different epitopes on the protein, and represented greater than 90% of the available epitopes. These epitopes were not unique to lupin nodule aspartate aminotransferase-P2 but were shown to be present on the enzyme from tobacco leaves and potato. Four of the epitopes were conformational with a fifth epitope recognized by the appropriate monoclonals in both its native and denatured forms. None of the monoclonal antibodies produced reacted with Rhizobium Iupini NZP2257 extracts. Antibodies against two epitopes showed some cross-reaction with the constitutive aspartate aminotransferase-P1 isoenzyme also found in lupin root nodules. However, affinity of these monoclonals for AAT-P1 was three orders of magnitude lower than for AAT-P2. Monoclonals against the other epitopes appeared to be specific for aspartate aminotransferase-P2.  相似文献   

17.
Pyruvate carboxylase was recently sequenced in Corynebacterium glutamicum and shown to play an important role of anaplerosis in the central carbon metabolism and amino acid synthesis of these bacteria. In this study we investigate the effect of the overexpression of the gene for pyruvate carboxylase (pyc) on the physiology of C. glutamicum ATCC 21253 and ATCC 21799 grown on defined media with two different carbon sources, glucose and lactate. In general, the physiological effects of pyc overexpression in Corynebacteria depend on the genetic background of the particular strain studied and are determined to a large extent by the interplay between pyruvate carboxylase and aspartate kinase activities. If the pyruvate carboxylase activity is not properly matched by the aspartate kinase activity, pyc overexpression results in growth enhancement instead of greater lysine production, despite its central role in anaplerosis and aspartic acid biosynthesis. Aspartate kinase regulation by lysine and threonine, pyruvate carboxylase inhibition by aspartate (shown in this study using permeabilized cells), as well as well-established activation of pyruvate carboxylase by lactate and acetyl coenzyme A are the key factors in determining the effect of pyc overexpression on Corynebacteria physiology.  相似文献   

18.
The kinetics of l-aspartate transport into pea chloroplasts was studied in the presence and absence of transport inhibitors to determine whether multiple aspartate carriers exist. Transport was measured by the silicone oil centrifugation technique. Reciprocal plots of concentration-dependent transport rates were biphasic, indicating the presence of two transport components, distinguishable on the basis of their affinity for aspartate. These transport components, called high affinity and low affinity transport could also be distinguished on the basis of their apparent substrate saturability and their sensitivity to media pH. The apparent Km for high affinity transport was 30 micromolar. The Km for low affinity transport was not determined. To test whether these transport components could also be distinguished on the basis of inhibitor sensitivity and to assess the value of inhibitors for distinguishing multiple aspartate translocators, a survey of several classes of potential inhibitors was conducted. High affinity aspartate transport was inhibited by p-chloromercuribenzenesulfonate and mersalyl, both sulfhydryl-reactive reagents; diethyl pyrocarbonate, a histidine-reactive reagent; and nigericin and carbonyl cyanide m-chlorophenylhydrazone, both ionophores. Low affinity aspartate transport was not inhibited by p-chloromercuribenzenesulfonate or nigericin, but preliminary results suggest it was sensitive to diethyl pyrocarbonate. Because the high and low affinity transport components could be distinguished not only by their sensitivity to media pH and substrate saturability, but also by their sensitivity to various inhibitors, we concluded that they may represent different transport systems or carriers.  相似文献   

19.
High specificity in aminoacylation of transfer RNAs (tRNAs) with the help of their cognate aminoacyl-tRNA synthetases (aaRSs) is a guarantee for accurate genetic translation. Structural and mechanistic peculiarities between the different tRNA/aaRS couples, suggest that aminoacylation systems are unrelated. However, occurrence of tRNA mischarging by non-cognate aaRSs reflects the relationship between such systems. In Saccharomyces cerevisiae, functional links between arginylation and aspartylation systems have been reported. In particular, it was found that an in vitro transcribed tRNAAsp is a very efficient substrate for ArgRS. In this study, the relationship of arginine and aspartate systems is further explored, based on the discovery of a fourth isoacceptor in the yeast genome, tRNA4Arg. This tRNA has a sequence strikingly similar to that of tRNAAsp but distinct from those of the other three arginine isoacceptors. After transplantation of the full set of aspartate identity elements into the four arginine isoacceptors, tRNA4Arg gains the highest aspartylation efficiency. Moreover, it is possible to convert tRNA4Arg into an aspartate acceptor, as efficient as tRNAAsp, by only two point mutations, C38 and G73, despite the absence of the major anticodon aspartate identity elements. Thus, cryptic aspartate identity elements are embedded within tRNA4Arg. The latent aspartate acceptor capacity in a contemporary tRNAArg leads to the proposal of an evolutionary link between tRNA4Arg and tRNAAsp genes.  相似文献   

20.
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