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 328 毫秒
1.
Aspartate transcarbamylase (EC 2.1.3.2) catalyzes the bi substrate reaction—carbamyl phosphate+ L-aspartate ? carbamyl aspartate ? phosphate, The order of addition of substrates and release of products for the homogeneous aspartate transcarbamylase fromPhaseolus aureuss eedlings has been investigated by using the kinetic methods of analysis. p ]Initial velocity studies indicated that the mechanism might be a sequential one. Product inhibition studies showed that phosphate was a linear competitive inhibitor with respect to carbamyl phosphate and was anS (slope) andI (intercept) linear noncompetitive inhibitor with respect to aspartate. Carbamyl aspartate was a noncompetitive inhibitor with respect to both the substrates. These inhibition patterns agreed with an ordered mechanism of reaction with carbamyl phosphate as the leading substrate and phosphate as the last product to leave the enzyme surface. The presence of dead end complexes and the rapid equilibrium random mechanism were ruled out by the absence of inhibition by the substrate(s) and the linear replot slopevs. the inhibitor concentration. Acetyl phosphate, an analog ue of carbamyl phosphate was a non-competitive inhibitor with respect to aspartate. This result could be explained both in terms of an ordered as well as a random mechanism. On the other hand, succinate, an analog ue of aspartate was an uncompetitive inhibitor with respect to carbamyl phosphate, indicating that the mechanism was ordered. p ]The transition state analog ue, N-(phosphonoacetyl)-L-aspartate, binds much more tightly than either of the two substrates. This analog ue was a linear competitive inhibitor with respect to carbamyl phosphate and a linear noncompetitive inhibitor with respect to aspartate. These results are compatible with an ordered mechanism rather than a random one.  相似文献   

2.
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   

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.
Isotopic exchange kinetics at equilibrium for E. coli native aspartate transcarbamylase at pH 7.8, 30 °C, are consistent with an ordered BiBi substrate binding mechanism. Carbamyl phosphate binds before l-Asp, and carbamyl-aspartate is released before inorganic phosphate. The rate of [14C]Asp C-Asp exchange is much faster than [32P]carbamyl phosphate Pi exchange. Phosphate, and perhaps carbamyl phosphate, appears to bind at a separate modifier site and prevent dissociation of active-site bound Pi or carbamyl phosphate. Initial velocity studies in the range of 0–40 °C reveal a biphasic Arrhenius plot for native enzyme: Ea (>15 °C) = 6.3 kcal/ mole and Ea (<15 °C) = 22.1 kcal/mole. Catalytic subunits show a monophasic plot with Ea ? 20.2 kcal/mole. This, with other data, suggests that with native enzyme a conformational change accompanying aspartate association contributes significantly to rate limitation at t > 15 °C, but that catalytic steps become definitively slower below 15 °C. Model kinetics are derived to show that this change in mechanism at low temperature can force an ordered substrate binding system to produce exchange-rate patterns consistent with a random binding system with all exchange rates equal. The nonlinear Arrhenius plot also has important consequences for current theories of catalytic and regulatory mechanisms for this enzyme.  相似文献   

5.
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   

6.
In contrast to holo-enzyme (c6r6), catalytic subunits (c3) of Escherichia coli aspartate transcarbamylase (carbamoyl-phosphate:L-aspartate carbamoyltransferase, EC 2.1.3.2) do not exhibit allosteric interactions or inhibition effects that complicate kinetic investigations of substrate binding order. Equilibrium isotope-exchange kinetic probes of c3 at pH 7.0 and 30 degrees C produced kinetic saturation patterns consistent with a strongly preferred order random kinetic mechanism, in which carbamoyl phosphate binds prior to aspartate and carbamoyl aspartate is released before Pi. Weak substrate inhibition effects observed with c6r6 did not occur with c3, possibly due to decreased affinity for ligands at the dianion inhibition site.  相似文献   

7.
Ornithine transcarbamylase (EC 2.1.3.3) was purified to homogeneity from rat liver. The basis of the method is the chromatography of a high-speed supernatant fraction of a homogenized rat liver on an affinity column consisting of the transition-state analog of ornithine transcarbamylase, δ-N-(phosphonacetyl)-l-ornithine, immobilized on epoxy-activated Sepharose 6B through the α-amino group. The enzyme was eluted from the column using a gradient of the substrate, carbamyl phosphate, and further purified by gel filtration. The enzyme elutes with a constant specific activity of 250 to 260 μmol min?1 mg?1 at pH 8.5, 37°C, and is free of contaminating proteins on sodium dodecyl sulfate gel electrophoresis. Determination of the molecular weight of the purified enzyme by centrifugation (98,000) and by gel electrophoresis in the presence of sodium dodecyl sulfate (35,300) indicates that the enzyme from rat liver is a trimer. The enzyme exhibits conventional Michaelis-Menten kinetics at pH 7.4 and in this respect differs from the enzyme prepared by other methods.  相似文献   

8.
The fluorescein dye, rose bengal in the dark: (i) inhibited the activity of mung bean aspartate transcarbamylase (EC 2.1.3.2) in a non-competitive manner, when aspartate was the varied substrate; (ii) induced a lag in the time course of reaction and this hysteresis was abolished upon preincubation with carbamyl phosphate; and (iii) converted the multiple bands observed on polyacrylamide gel electrophoresis of enzyme into a single band. The binding of the dye to the enzyme induced a red shift in the visible spectrum of dye suggesting that it was probably interacting at a hydrophobic region in the enzyme. The dye, in the presence of light, inactivated the enzyme and the inactivation was not dependent on pH. All the effects of the dye could be reversed by UMP, an allosteric inhibitor of the enzyme. The loss of enzyme activity on photoinactivation and the partial protection afforded by N-phosphonoacetyl-L-aspartate, a transition state analog and carbamyl phosphate plus succinate, a competitive inhibitor for aspartate, as well as the reversal of the dye difference spectrum by N-phosphonoacetyl-L-aspartate suggested that in the mung bean aspartate transcarbamylase, unlike in the case ofEscherichia coli enzyme, the active and allosteric sites may be located close to each other.  相似文献   

9.
Aspartate transcarbamylase (EC 2·1·3·2) purified from mung bean seedlings was used as a model to understand the mechanism of allosteric regulation. The enzyme exhibited homotropic interactions with carbamyl phosphate. Preincubation of the enzyme with aspartate abolished the sigmoidicity of the carbamyl phosphate saturation curve. UMP was the most potent inhibitor of the reaction and was noncompetitive with respect to aspartate. The sigmoidicity of carbamyl phosphate saturation curves increased with increase in UMP concentration. These results were analysed by an iterative least squares procedure. There was no change inV max values with increase in the UMP concentration, although theK 0·5 values (concentration of carbamyl phosphate required to reach half maximal velocity) increased. This implied that the effect of UMP was on the binding of carbamyl phosphate only and not on the catalytic function of the enzyme. The allosteric properties of the enzyme could be explained in terms ofK system of the symmetry model. The values of the allosteric constantsn, L andc calculated for mung bean enzyme, making use of the Monod equation accounted for all the observed properties. The enzyme appeared to be a tetramer (n=4) and in the absence of ligands was predominantly in theT form (L o= 2·25). Carbamyl phosphate bound preferentially to theR form (c= 10?3), while UMP bound preferentially to theT form and hence these two ligands exhibited the typical heterotropic interactions as expected of antagonistic ligands.  相似文献   

10.
Drosophila cells were treated in vitro with N-phosphonacetyl- -aspartate (PALA) which is a specific inhibitor of aspartate transcarbamylase, the second enzyme of the pyrimidine biosynthetic pathway. By stepwise selection using increasing amounts of this inhibitor, PALA-resistant (PALAr) stable clones have been isolated. Enzymatic activities of aspartate transcarbamylase, carbamyl phosphate synthetase and dihydro-orotase, borne by the same multifunctional protein, CAD, are increased 6–12-fold in these resistant clones compared with parental cells. The aspartate transcarbamylase in PALAr cells is shown by physical, kinetic and immunological criteria to be normal. The data from immunotitration and immunoblotting experiments indicate that the increased enzyme activities result from the overproduction of CAD.  相似文献   

11.
Aspartate transcarbamylase (carbamoyl-phosphate: L-aspartate carbamoyltransferase, EC 2.1.3.2) has been purified from Mycobacterium smegmatis TMC 1546 using streptomycin sulphate precipitation, ammonium sulphate precipitation, DE-52 chromatography, second ammonium sulphate precipitation, Sephadex G-200 gel filtration, and aspartate-linked CNBr-activated Sepharose 4B affinity chromatography in successive order. The enzyme was purified 231.6-fold, and the preparation was found to be homogeneous on column chromatography and polyacrylamide gel electrophoresis. The purified enzyme had a molecular weight of 246,000 and was composed of two asymmetrical subunits. The kinetic and regulatory properties of aspartate transcarbamylase from M. smegmatis were also studied. The enzyme was found to be an allosteric in nature with carbamyl phosphate showing positive cooperativity and UMP exhibiting a negative cooperativity. CTP was found to be the most potent inhibitor among nucleotides. Phosphate acted as a non-competitive product inhibitor with respect to aspartate. Succinate and maleate exerted a competitive inhibition when aspartate was the variable substrate.  相似文献   

12.
We have characterized further the biogenesis in vitro of ornithine transcarbamylase, a homotrimeric mitochondrial matrix enzyme synthesized in the cytoplasm as a larger precursor. When cell-free translation mixtures containing the ornithine transcarbamylase precursor (40 kDa) were chromatographed on Bio-Gel P-200 columns, all of the precursor eluted as aggregates or complexes with molecular weights greater than 200 kDa. None of the precursor bound to a ligand affinity column containing delta-N-(phosphonoacetyl)-L-ornithine (delta-PALO), a transition-state analog and competitive inhibitor of carbamyl phosphate binding, which recognizes native ornithine transcarbamylase. In contrast, a significant portion of the labeled mature-sized subunits, formed when intact mitochondria processed the precursor, bound specifically to the delta-PALO column, were eluted by carbamyl phosphate, and chromatographed on a Bio-Gel P-300 column with a mobility identical to that of native, trimeric ornithine transcarbamylase. No such binding to delta-PALO was observed for the mature-sized monomer or dimer, or for the intermediate-sized ornithine transcarbamylase polypeptide. Moreover, processing by a mitochondrial matrix fraction failed to yield trimeric enzyme, despite producing ample amounts of mature-sized monomer. We conclude that delta-PALO recognizes only trimeric ornithine transcarbamylase composed of mature-sized subunits and that such trimers can be assembled in vitro by intact mitochondria following translocation and proteolytic processing.  相似文献   

13.
A steady-state kinetic analysis has been performed on the pyruvate dehydrogenase complex from pea (Pisum sativum L.) mitochondria and castor bean (Ricinus communis L.) proplastids. Substrate interaction kinetics for all substrates gave parallel lines consistent with a multisite ping-pong mechanism. Product inhibition studies showed uncompetitive inhibition between acetyl-CoA and pyruvate and competitive inhibition between NADH and NAD+, both of which are also consistent with this mechanism. In the mitochondrial complex, acetyl-CoA showed noncompetitive inhibition versus CoA which suggests that the intermediate complex is kinetically important in the lipoamide transacetylase component of this complex. In contrast, the proplastid complex showed competitive inhibition in this interaction. NADH is a noncompetitive inhibitor versus CoA in both complexes indicating that these complexes, like the mammalian complex, may have protein-protein interactions between the second and third enzymes of the complex. Since NADH also shows noncompetitive inhibition versus pyruvate, this interaction may extend to all components of the complex. Acetyl-CoA shows noncompetitive inhibition versus NAD+ which may also be a result of interaction between the second and third enzymes of the complex. The limiting Michaelis constants for substrates and the inhibitor constants for both complexes were determined.  相似文献   

14.
Glutamine-dependent carbamyl phosphate synthetase of rat liver, purified about 2,100-fold, existed as a complex with aspartate transcarbamylase and dihydroorotase, the second and third enzymes of pyrimidine biosynthesis, with a sedimentation coefficient of 27 S. Treatment of this complex with pancreatic elastase caused a selective inactivation of the transcarbamylase with concomitant dissociation of the complex. The dissociated synthetase was as sensitive to allosteric effectors as the enzyme within the complex, but had a 5 times higher apparent Km for MgATP2?. This change appears to be intimately related to the release of the enzyme from the complex.  相似文献   

15.
Mammalian and hepatic aspartate transcarbamylase is inhibited by phenobarbital p-nitrophenylhydra-zone in a reversible and non-competitive type with Ki values 8.45 × 10?5 and 9.64×10?5 M in the reactions toward carbamyl phosphate and aspartate, respectively. In vivo inhibition occurred in a dose-dependent manner in which less than 50% of the activity was retained. These observations suggest that this inhibitor may interfere with the in vivo regulation of this enzyme and lead to an additional biological effect of phenobarbitals.  相似文献   

16.
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   

17.
Pyrimidine-repressible carbamyl phosphate synthetase P was synthesized in parallel with aspartate transcarbamylase during growth of Bacillus subtilis on glucose-nutrient broth. Both enzymes were inactivated at the end of exponential growth, but at different rates and by different mechanisms. Unlike the inactivation of aspartate transcarbamylase, the inactivation of carbamyl phosphate synthetase P was not interrupted by deprivation for oxygen or in a tricarboxylic acid cycle mutant. The arginine-repressible isozyme carbamyl phosphate synthetase A was synthesized in parallel with ornithine transcarbamylase during the stationary phase under these growth conditions. Again, both enzymes were subsequently inactivated, but at different rates and by apparently different mechanisms. The inactivation of carbamyl phosphate synthetase A was not affected in a protease-deficient mutatn the inactivation of ornithine transcarbamylase was greatly slowed.  相似文献   

18.
The kinetic properties of wheat germ sucrose phosphate synthetase, which catalyzes the reaction UDP-glucose + fructose 6-phosphate → UDP + sucrose 6-phosphate have been studied. A plot of the reciprocal initial velocity versus reciprocal substrate concentration gave a series of intersecting lines indicating a sequential mechanism. Product inhibition studies showed that UDP was competitive with UDP-glucose and noncompetitive with fructose 6-phosphate. A dead-end inhibitor, inorganic phosphate, was competitive with UDP-glucose and noncompetitive with fructose 6-phosphate. The results of initial velocity and product and dead-end inhibition studies suggested that the addition of substrates to the enzyme follows an ordered mechanism.  相似文献   

19.
Transcarbamylases catalyze the transfer of the carbamyl group from carbamyl phosphate (CP) to an amino group of a second substrate such as aspartate, ornithine, or putrescine. Previously, structural determination of a transcarbamylase from Xanthomonas campestris led to the discovery of a novel N-acetylornithine transcarbamylase (AOTCase) that catalyzes the carbamylation of N-acetylornithine. Recently, a novel N-succinylornithine transcarbamylase (SOTCase) from Bacteroides fragilis was identified. Structural comparisons of AOTCase from X. campestris and SOTCase from B. fragilis revealed that residue Glu92 (X. campestris numbering) plays a critical role in distinguishing AOTCase from SOTCase. Enzymatic assays of E92P, E92S, E92V, and E92A mutants of AOTCase demonstrate that each of these mutations converts the AOTCase to an SOTCase. Similarly, the P90E mutation in B. fragilis SOTCase (equivalent to E92 in X. campestris AOTCase) converts the SOTCase to AOTCase. Hence, a single amino acid substitution is sufficient to swap the substrate specificities of AOTCase and SOTCase. X-ray crystal structures of these mutants in complexes with CP and N-acetyl-L-norvaline (an analog of N-acetyl-L-ornithine) or N-succinyl-L-norvaline (an analog of N-succinyl-L-ornithine) substantiate this conversion. In addition to Glu92 (X. campestris numbering), other residues such as Asn185 and Lys30 in AOTCase, which are involved in binding substrates through bridging water molecules, help to define the substrate specificity of AOTCase. These results provide the correct annotation (AOTCase or SOTCase) for a set of the transcarbamylase-like proteins that have been erroneously annotated as ornithine transcarbamylase (OTCase, EC 2.1.3.3).  相似文献   

20.
The search of an index for the instantaneous estimate of the in situ growth rate of marine animals led us to attempt to measure the specific activity of aspartate transcarbamylase (ATC). The experiments to test the value of the index were carried out on the scallop Pecten maximus L. p]The first step was to find the optimum conditions for enzyme activity measurement. At 35 °C. the scallop ATC shows an optimum pH of 9 and a Km of 4.6 × 10?3 M for aspartate and of 8.0 × 10?4 M for carbamyl phosphate. The different types of inhibition by the substrates high concentrations and the products suggest an ordered sequential mechanism for the reaction. The decrease in enzyme activity due to metallic ions (Cu2+ and Zn2+) and to parahydroxymercuribenzoate is compatible with the presence of a sulphydryl group in the active site. p]The variations in ATC levels within the gonad and within the mantle of the scallop were measured and compared with the processes of the sexual maturation and somatic growth in a natural population. For the two tissues, a correlation between the ATC specific activity and the relative growth rate is demonstrated.  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