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.
13C and 15N isotope effects have been measured for the aspartate transcarbamylase (ATCase) reaction in an effort to elucidate the chemical mechanism of this highly regulated enzyme. The observed 15(V/K(asp))H2O value for the ATCase holoenzyme at saturing carbamyl phosphate and 12 mM L-aspartate is 1.0045 at pH 7.5, and this value remains unchanged in the presence of 5 mM ATP (activator) or 5 mM CTP (inhibitor). The fact that the isotope effect is not changed by the allosteric modifiers supports the conclusion that the kinetic properties of the active form of ATCase are not influenced by ATP or CTP. The observed 15(V/K(asp)) values for the catalytic subunit of ATCase are also the same as those determined for the holoenzyme, suggesting that the chemical mechanisms of both enzyme species are the same. Quantitative analysis of 13C and 15N isotope effects in both H2O and D2O has led to the proposal of a chemical model for the ATCase reaction which involves a precatalytic conformational change to form an activated complex that facilitates deprotonation of L-aspartate by an enzyme functional group. Nucleophilic attack on the carbonyl carbon of carbamyl phosphate by the alpha-amino group of L-aspartate results in the formation of a tetrahedral intermediate. An intramolecular proton transfer leads to formation of products N-carbamyl-L-aspartate and inorganic phosphate.  相似文献   

2.
13C isotope effects have been measured for the aspartate transcarbamylase holoenzyme (ATCase) and catalytic subunit catalyzed reactions in the presence of the bisubstrate analog N-(phosphonoacetyl)-L-aspartate (PALA). For holoenzyme-catalyzed reactions in the physiological direction with very low levels of L-aspartate as substrate, or with L-cysteine sulfinate as substrate, or in the reverse direction with carbamyl-L-aspartate and phosphate as substrates, the isotope effect data show a slight dependence on PALA concentration. Under these conditions, PALA first stimulates the rate and then inhibits it at higher concentrations. The observed isotope effect at maximum stimulation by PALA is slightly smaller than in the absence of the analog, but as the PALA concentration is increased to reduce the rate to its original value, the observed isotope effect also increases and approaches the value of the isotope effect determined in the absence of PALA. These data suggest that the kinetic properties of the active enzyme are affected by the number of active sites occupied by PALA, indicating communication between subunits, and a mathematical model is proposed which explains our experimental observations. In contrast to these results with the holoenzyme, isotope effects measured for the reaction catalyzed by the isolated catalytic subunits are not altered in the presence of PALA. Taken together, these data are consistent with the two-state model for the homotropic regulation of ATCase.  相似文献   

3.
The allosteric transition of Escherichia coli aspartate transcarbamylase involves significant alterations in structure at both the quaternary and tertiary levels. On the tertiary level, the 240s loop (residues 230-245 of the catalytic chain) repositions, influencing the conformation of Arg-229, a residue near the aspartate binding site. In the T state, Arg-229 is bent out of the active site and may be stabilized in this position by an interaction with Glu-272. In the R state, the conformation of Arg-229 changes, allowing it to interact with the beta-carboxylate of aspartate, and is stabilized in this position by a specific interaction with Glu-233. In order to ascertain the function of Arg-229, Glu-233, and Glu-272 in the catalytic and cooperative interactions of the enzyme, three mutant enzymes were created by site-specific mutagenesis. Arg-229 was replaced by Ala, while both Glu-233 and Glu-272 were replaced by Ser. The Arg-229----Ala and Glu-233----Ser enzymes exhibit 10,000-fold and 80-fold decreases in maximal activity, respectively, and they both exhibit a 2-fold increase in the aspartate concentration at half the maximal observed velocity, [S]0.5. The Arg-229----Ala enzyme still exhibits substantial homotropic cooperativity, but all cooperativity is lost in the Glu-233----Ser enzyme. The Glu-233----Ser enzyme also shows a 4-fold decrease in the carbamyl phosphate [S]0.5, while the Arg-229----Ala enzyme shows no change in the carbamyl phosphate [S]0.5 compared to the wild-type enzyme. The Glu-272 to Ser mutation results in a slight reduction in maximal activity, an increase in [S]0.5 for both aspartate and carbamyl phosphate, and reduced cooperativity. Analysis of the isolated catalytic subunits from these three mutant enzymes reveals that in each case the changes in the kinetic properties of the isolated catalytic subunit are similar to the changes caused by the mutation in the holoenzyme. PALA was able to activate the Glu-233----Ser enzyme, at low aspartate concentrations, even though the mutant holoenzyme did not exhibit any cooperativity, indicating that cooperative interactions still exist between the active sites in this enzyme. It is proposed that Glu-233 of the 240s loop helps create the high-activity-high-affinity R state by positioning the side chain of Arg-229 for aspartate binding while Glu-272 helps stabilize the low-activity-low-affinity T state by positioning the side chain of Arg-229 so that it cannot interact with aspartate.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

4.
Two mutant versions of Escherichia coli aspartate transcarbamylase were created by site-specific mutagenesis. Arg-234 of the 240s loop was replaced by serine in order to help deduce the function of the interactions that normally occur between Arg-234 and both Glu-50 and Gln-231 in the R state of the enzyme. The other mutation involved the replacement of Asp-271 by asparagine to further test the functional importance of the Tyr-240-Asp-271 link that has previously been proposed to stabilize the T state of the enzyme [Middleton, S. A., & Kantrowitz, E. R. (1986) Proc. Natl. Acad. Sci. U.S.A. 83, 5866-5870]. The Arg-234----Ser holoenzyme exhibits no cooperativity, a 24-fold reduction in maximal velocity, normal affinity for carbamyl phosphate, and substantially reduced affinity for aspartate and N-(phosphonoacetyl)-L-aspartate (PALA). Unlike the wild-type enzyme, the heterotropic effectors ATP and CTP are able to influence the activity of the Arg-234----Ser enzyme at saturating aspartate concentrations. The Arg-234----Ser catalytic subunit exhibits a 33-fold reduction in maximal activity, an aspartate Km of 261 mM, compared to 5.7 mM for the wild-type catalytic subunit, and only a small alteration in the Km for carbamyl phosphate. Together these results provide additional evidence that the interdomain bridging interactions between Glu-50 of the carbamyl phosphate domain and both Arg-167 and Arg-234 of the aspartate domain are necessary for the stabilization of the high-activity-high-affinity configuration of the active site of the enzyme. Furthermore, without the interdomain bridging interactions, the holoenzyme no longer exhibits homotropic cooperativity.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
Most investigations of the allosteric properties of the regulatory enzyme aspartate transcarbamoylase (ATCase) from Escherichia coli are based on the sigmoidal dependence of enzyme activity on substrate concentration and the effects of the inhibitor, CTP, and the activator, ATP, on the saturation curves. Interpretations of these effects in terms of molecular models are complicated by the inability to distinguish between changes in substrate binding and catalytic turnover accompanying the allosteric transition. In an effort to eliminate this ambiguity, the binding of the 3H-labeled bisubstrate analog N-(phosphonacetyl)-L-aspartate (PALA) to aspartate transcarbamoylase in the absence and presence of the allosteric effectors ATP and CTP has been measured directly by equilibrium dialysis at pH 7 in phosphate buffer. PALA binds with marked cooperativity to the holoenzyme with an average dissociation constant of 110 nM. ATP and CTP alter both the average affinity of ATCase for PALA and the degree of cooperativity in the binding process in a manner analogous to their effects on the kinetic properties of the enzyme; the average dissociation constant of PALA decreases to 65 nM in the presence of ATP and increases to 266 nM in the presence of CTP while the Hill coefficient, which is 1.95 in the absence of effectors, becomes 1.35 and 2.27 in the presence of ATP and CTP, respectively. The isolated catalytic subunit of ATCase, which lacks the cooperative kinetic properties of the holoenzyme, exhibits only a very slight degree of cooperativity in binding PALA. The dissociation constant of PALA from the catalytic subunit is 95 nM. Interpretation of these results in terms of a thermodynamic scheme linking PALA binding to the assembly of ATCase from catalytic and regulatory subunits demonstrates that saturation of the enzyme with PALA shifts the equilibrium between holoenzyme and subunits slightly toward dissociation. Ligation of the regulatory subunits by either of the allosteric effectors leads to a change in the effect of PALA on the association-dissociation equilibrium.  相似文献   

6.
We present evidence that cysteine 269 of the small subunit of Escherichia coli carbamyl phosphate synthetase is essential for the hydrolysis of glutamine. When cysteine 269 is replaced with glycine or with serine by site-directed mutagenesis of the carA gene, the resulting enzymes are unable to catalyze carbamyl phosphate synthesis with glutamine as nitrogen donor. Even though the glycine 269, and particularly the serine 269 enzyme bind significant amounts of glutamine, neither glycine 269 nor serine 269 can hydrolyze glutamine. The mutations at cysteine 269 do not affect carbamyl phosphate synthesis with NH3 as substrate. The NH3-dependent activity of the mutant enzymes was equal to that of wild-type. Measurements of Km indicate that the enzyme uses unionized NH3 rather than ammonium ion as substrate. The apparent Km for NH3 of the wild-type enzyme is calculated to be about 5 mM, independent of pH. The substitution of cysteine 269 with glycine or with serine results in a decrease of the apparent Km value for NH3 from 5 mM with the wild-type to 3.9 mM with the glycine, and 2.9 mM with the serine enzyme. Neither the glycine nor the serine mutation at position 269 affects the ability of the enzyme to catalyze ATP synthesis from ADP and carbamyl phosphate. Allosteric properties of the large subunit are also unaffected. However, substitution of cysteine 269 with glycine or with serine causes an 8- and 18-fold stimulation of HCO-3 -dependent ATPase activity, respectively. The increase in ATPase activity and the decrease in apparent Km for NH3 provide additional evidence for an interaction of the glutamine binding domain of the small subunit with one of the two known ATP sites of the large subunit.  相似文献   

7.
Interaction between a 70-amino acid and zinc-binding polypeptide from the regulatory chain and the catalytic (C) trimer of aspartate transcarbamoylase (ATCase) leads to dramatic changes in enzyme activity and affinity for active site ligands. The hypothesis that the complex between a C trimer and 3 polypeptide fragments (zinc domain) is an analog of R state ATCase has been examined by steady-state kinetics, heavy-atom isotope effects, and isotope trapping experiments. Inhibition by the bisubstrate ligand, N-(phosphonacetyl)-L-aspartate (PALA), or the substrate analog, succinate, at varying concentrations of substrates, aspartate, or carbamoyl phosphate indicated a compulsory ordered kinetic mechanism with carbamoyl phosphate binding prior to aspartate. In contrast, inhibition studies on C trimer were consistent with a preferred order mechanism. Similarly, 13C kinetic isotope effects in carbamoyl phosphate at infinite aspartate indicated a partially random kinetic mechanism for C trimer, whereas results for the complex of C trimer and zinc domain were consistent with a compulsory ordered mechanism of substrate binding. The dependence of isotope effect on aspartate concentration observed for the Zn domain-C trimer complex was similar to that obtained earlier for intact ATCase. Isotope trapping experiments showed that the compulsory ordered mechanism for the complex was attributable to increased "stickiness" of carbamoyl phosphate to the Zn domain-C trimer complex as compared to C trimer alone. The rate of dissociation of carbamoyl phosphate from the Zn domain-C trimer complex was about 10(-2) that from C trimer.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
The substitution of alanine for lysine at position 56 of the regulatory polypeptide of aspartate transcarbamoylase affected both homotropic and heterotropic characteristics. In the absence of effectors, the ALAr56-substituted holoenzyme lost the homotropic cooperativity observed for aspartate in the wild-type holoenzyme. Under conditions of allosteric inhibition in the presence of 2mM CTP, the cooperative character of ATCase was restored, and the Hill coefficient increased from 1.0 to 1.7. In contrast to the native enzyme, the altered enzyme did not respond to ATP; however, ATP could still bind to the enzyme as demonstrated by its direct competition with CTP. Furthermore, the recently observed CTP-UTP synergism of the wild-type enzyme was not detectable. The site-directed mutant enzyme could not be activated by low levels of the bisubstrate analogue, N-(phosphonacetyl)-L-aspartate, and the rate of association of pHMB with the cysteine residues located at the interface of the catalytic and regulatory chains was slightly altered. These characteristics suggested that the mutant holoenzyme assumed a relaxed (or abnormal T state) conformation. Thus, this single substitution differentially affected the heterotropic responses to the various allosteric effectors of ATCase and eliminated the homotropic characteristics in response to aspartate in the absence of CTP.  相似文献   

9.
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   

10.
11.
Saturation curves of activity versus concentration were determined for aspartate transcarbamylase from Escherichia coli (EC 2.1.3.2) for the substrate L-aspartate at saturating carbamyl phosphate (4.8 mM) in buffered solution at pH values from 6.0 to 12.0. Hill coefficients were obtained from the sigmoidal curves. At pH values from 7.8 to 9.1, where substrate inhibition causes difficulties in the Hill approximation, our kinetic scheme includes substrate inhibition and residual activity in the abortive enzyme-substrate complex. The plot of Hill coefficient versus pH has pKalpha values of 7.4 and 9.8 at the half-maximum positions of the curve which has a plateau from pH 8.1 to 9.1. These pKalpha values may be associated with functional groups involved in the allosteric transition which activates the enzyme. A plot of [S]0.5 versus pH shows a pKalpha of 8.5, which may belong to a residue either at or near the aspartate binding site. At 50 mM aspartate concentration the pH-rate profile shows maxima at pH values of 8.8 and 10.0 (cf. Weitzman, P.D.J., and Wilson, I.B.(1966)J. Biol. Chem. 2418 5481-5488, who used 100 mM aspartate). However, when the pH-dependent substrate inhibition is included, the calculated Vmax--H curve is bell-shaped like that of the isolated catalytic subunit.  相似文献   

12.
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   

13.
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   

14.
A detailed kinetic analysis of the catalytic trimer of aspartate transcarbamoylase containing the active site substitution H134A was performed to investigate the role of His 134 in the catalytic mechanism. Replacement of histidine by alanine resulted in decreases in the affinities for the two substrates, carbamoyl phosphate and aspartate, and the inhibitor succinate, by factors of 50, 10, and 6, respectively, and yielded a maximum velocity that was 5% that of the wild-type enzyme. However, the pK values determined from the pH dependence of the kinetic parameters, log V and log (V/K) for aspartate, the pK(i) for succinate, and the pK(ia) for carbamoyl phosphate, were similar for both the mutant and the wild-type enzymes, indicating that the protonated form of His 134 does not participate in binding and catalysis between pH 6.2 and 9.2. 13C and 15N isotope effects were studied to determine which steps in the catalytic mechanism were altered by the amino acid substitutions. The 13(V/K) for carbamoyl phosphate exhibited by the catalytic trimer containing alanine at position 134 revealed an isotope effect of 4.1%, probably equal to the intrinsic value and, together with quantitative analysis of the 15N isotope effects, showed that formation of the tetrahedral intermediate is rate-determining for the mutant enzyme. Thus, His 134 plays a role in the chemistry of the reaction in addition to substrate binding. The initial velocity pattern for the reaction catalyzed by the H134A mutant intersected to the left of the vertical axis, negating an equilibrium ordered kinetic mechanism.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
Y J Farrar  G M Carlson 《Biochemistry》1991,30(42):10274-10279
The phosphorylase kinase holoenzyme from skeletal muscle is composed of a catalytic and three different regulatory subunits. Analysis of the kinetic mechanism of the holoenzyme is complicated because both the natural substrate phosphorylase b and also phosphorylase kinase itself have allosteric binding sites for adenine nucleotides. In the case of the kinase, these allosteric sites are not on the catalytic subunit. We have investigated the kinetic mechanism of phosphorylase kinase by using its isolated catalytic gamma-subunit (activated by calmodulin) and an alternative peptide substrate (SDQEKRKQISVRGL) corresponding to the convertible region of phosphorylase b, thus eliminating from our system all known allosteric binding sites for nucleotides. This peptide has been previously employed to study the kinetic mechanism of the kinase holoenzyme before the existence of the allosteric sites on the regulatory subunits was suspected [Tabatabai, L. B., & Graves, D. J. (1978) J. Biol. Chem. 253, 2196-2202]. This peptide was determined to be as good an alternative substrate for the isolated catalytic subunit as it was for the holoenzyme. Initial velocity data indicated a sequential kinetic mechanism with apparent Km's for MgATP and peptide of 0.07 and 0.47 mM, respectively. MgADP used as product inhibitor showed competitive inhibition against MgATP and noncompetitive inhibition against peptide, whereas with phosphopeptide as product inhibitor, the inhibition was competitive against both MgATP and peptide. The initial velocity and product inhibition studies were consistent with a rapid equilibrium random mechanism with one abortive complex, enzyme-MgADP-peptide. The substrate-directed, dead-end inhibitors 5'-adenylyl imidodiphosphate and Asp-peptide, in which the convertible Ser of the alternative peptide substrate was replaced with Asp, were competitive inhibitors toward their like substrates and noncompetitive inhibitors toward their unlike substrates, further supporting a random mechanism, which was also the conclusion from the report cited above that used the holoenzyme.  相似文献   

16.
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   

17.
The aspartate transcarbamoylases (ATCase, EC 2.1.3.2) of Escherichia coli and Serratia marcescens have similar dodecameric enzyme structures (2(c3):3(r2] but differ in both regulatory and catalytic characteristics. The catalytic cistrons (pyrB) of the ATCases from E. coli and S. marcescens encode polypeptides of 311 and 306 amino acids, respectively; there is a 76% identity between the DNA sequences and an overall amino acid homology of 88% (38 differences). The regulatory cistrons (pyrI) of these ATCases encode polypeptides of 153 and 154 amino acids, respectively, and there is a 75% identity between the DNA sequences and an overall amino acid homology of 77% (36 differences). In both species, the two genes are arranged as a bicistronic operon, with pyrB promoter proximal. A comparison of the deduced amino acid sequences reveals that the active site and the allosteric binding sites, as well as most of the intrasubunit interactions and intersubunit associations, are conserved in the E. coli and the S. marcescens enzymes; however, there are specific differences which undoubtedly contribute to the catalytic and regulatory differences between the enzymes of the two species. These differences include residues that have been implicated in the T-R transition, c1:r1 interface interactions, and the CTP binding site. A hybrid ATCase assembled in vivo with catalytic subunits from E. coli and regulatory subunits from S. marcescens has a 6 mM requirement for aspartate at half-maximal saturation, similar to the 5.5 mM aspartate requirement of the native E. coli holoenzyme at half-maximal saturation. However, the heterotropic response of this hybrid enzyme is characteristic of the heterotropic response of the native S. marcescens holoenzyme: ATP activation and CTP activation. Activation by both allosteric effectors indicates that the heterotropic response of this hybrid holoenzyme (Cec:Rsm) is determined by the associated S. marcescens regulatory subunits.  相似文献   

18.
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   

19.
In the catalytic chain of Escherichia coli aspartate transcarbamylase, Tyr240 helps stabilize the T-state conformation by an intrachain hydrogen bond to Asp271. Changes in kinetic characteristics of ATCase that result from disruption of this bond by site-specific mutation of Tyr240----Phe have been investigated by isotopic exchanges at chemical equilibrium. The Tyr240----Phe (Y240F) mutation caused the rate of the [32P] carbamyl phosphate (C-P) in equilibrium Pi exchange to decrease by 2-8-fold, without altering the [14C]Asp in equilibrium N-carbamyl-L-aspartate (C-Asp) rate. The mutation also caused the S0.5 and Hill nH values to decrease in virtually every substrate saturation experiment. Upon increasing the concentrations of the C-P,Pi or C-P,C-Asp reactant-product pairs, inhibition effects observed with the C-P in equilibrium Pi exchange for wild-type enzyme were not apparent with the Y240F mutant enzyme. In contrast, upon increasing the concentrations of the Asp,C-Asp and Asp,Pi pairs, inhibition effects on C-P in equilibrium Pi observed with wild-type enzyme became stronger with the Y240F mutant enzyme. These data indicate that the Tyr240----Phe mutation alters the kinetic mechanism in two different ways: on the reactant side, C-P binding prior to Asp shifts from preferred to compulsory order, and, on the product side, C-Asp and Pi release changes from preferred to nearly random order. These conclusions were also confirmed on a quantitative basis by computer simulations and fitting of the data, which also produced an optimal set of rate constants for the Y240F enzyme. The Arrhenius plot for wild-type holoenzyme was biphasic, but those for catalytic subunits and Y240F enzyme were linear (monophasic). Taken together, the data indicate that the Tyr240----Phe mutation destabilizes the T-state and shifts the equilibrium for the T-R allosteric transition toward the R-state by increasing the rate of T----R conversion.  相似文献   

20.
Here we report the first use of disulfide bond formation to stabilize the R allosteric structure of Escherichia coli aspartate transcarbamoylase. In the R allosteric state, residues in the 240s loop from two catalytic chains of different subunits are close together, whereas in the T allosteric state they are far apart. By substitution of Ala-241 in the 240s loop of the catalytic chain with cysteine, a disulfide bond was formed between two catalytic chains of different subunits. The cross-linked enzyme did not exhibit cooperativity for aspartate. The maximal velocity was increased, and the concentration of aspartate required to obtain one-half the maximal velocity, [Asp](0.5), was reduced substantially. Furthermore, the allosteric effectors ATP and CTP did not alter the activity of the cross-linked enzyme. When the disulfide bonds were reduced by the addition of 1,4-dithio-dl-threitol the resulting enzyme had kinetic parameters very similar to those observed for the wild-type enzyme and regained the ability to be activated by ATP and inhibited by CTP. Small-angle x-ray scattering was used to verify that the cross-linked enzyme was structurally locked in the R state and that this enzyme after reduction with 1,4-dithio-dl-threitol could undergo an allosteric transition similar to that of the wild-type enzyme. The complete abolition of homotropic and heterotropic regulation from stabilizing the 240s loop in its closed position in the R state, which forms the catalytically competent active site, demonstrates the significance that the quaternary structural change and closure of the 240s loop has in the functional mechanism of aspartate transcarbamoylase.  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