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 40 毫秒
1.
Huang J  Lipscomb WN 《Biochemistry》2006,45(2):346-352
An X-ray diffraction study to 2.0 A resolution shows that this enzyme, ATCase, is in the T-state (the c3 to c3 distance is 45.2 A) when ATCase is bound to carbamyl phosphate (CP) and to L-alanosine (an analogue of aspartate). This result strongly supports the kinetic results that alanosine did not inhibit the carbamylation of aspartate in the normal reaction of native ATCase plus CP and aspartate [Baillon, J., Tauc, P., and Hervé, G. (1985) Biochemistry 24, 7182-7187]. The structure further reveals that the phosphate of CP is 4 A away from its known position in the R-state and is in the T-state position of P(i) in a recent study of ATCase complexed with products, phosphate (P(i)) and N-carbamyl-L-aspartate [Huang, J., and Lipscomb, W. N. (2004) Biochemistry 43, 6422-6426]. Moreover, the alanosine position in this T-state is somewhat displaced from that expected for its analogue, aspartate, from the R-state position. The relations of these structural aspects to the kinetics are presented.  相似文献   

2.
Van Boxstael S  Maes D  Cunin R 《The FEBS journal》2005,272(11):2670-2683
Aspartate transcarbamylase (ATCase) (EC 2.1.3.2) from the hyperthermophilic archaeon Pyrococcus abyssi was purified from recombinant Escherichia coli cells. The enzyme has the molecular organization of class B microbial aspartate transcarbamylases whose prototype is the E. coli enzyme. P. abyssi ATCase is cooperative towards aspartate. Despite constraints imposed by adaptation to high temperature, the transition between T- and R-states involves significant changes in the quaternary structure, which were detected by analytical ultracentrifugation. The enzyme is allosterically regulated by ATP (activator) and by CTP and UTP (inhibitors). Nucleotide competition experiments showed that these effectors compete for the same sites. At least two regulatory properties distinguish P. abyssi ATCase from E. coli ATCase: (a) UTP by itself is an inhibitor; (b) whereas ATP and UTP act at millimolar concentrations, CTP inhibits at micromolar concentrations, suggesting that in P. abyssi, inhibition by CTP is the major control of enzyme activity. While V(max) increased with temperature, cooperative and allosteric effects were little or not affected, showing that molecular adaptation to high temperature allows the flexibility required to form the appropriate networks of interactions. In contrast to the same enzyme in P. abyssi cellular extracts, the pure enzyme is inhibited by the carbamyl phosphate analogue phosphonacetate; this difference supports the idea that in native cells ATCase interacts with carbamyl phosphate synthetase to channel the highly thermolabile carbamyl phosphate.  相似文献   

3.
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   

4.
Comparative modeling of mammalian aspartate transcarbamylase   总被引:1,自引:0,他引:1  
J L Scully  D R Evans 《Proteins》1991,9(3):191-206
Mammalian aspartate transcarbamylase (ATCase) is part of a 243 kDa multidomain polypeptide, called CAD, that catalyzes the first three steps in de novo pyrimidine biosynthesis. The structural organization of the mammalian enzyme is very different from E. coli ATCase, a dodecameric, monofunctional molecule comprised of six copies of separate catalytic and regulatory chains. Nevertheless, sequence similarities and other properties suggested that the mammalian ATCase domain and the E. coli ATCase catalytic chain have the same tertiary fold. A model of mammalian ATCase was built using the X-ray coordinates of the E. coli catalytic chain as a tertiary template. Five small insertions and deletions could be readily accommodated in the model structure. Following energy minimization the RMS difference in the alpha carbon positions of the mammalian and bacterial proteins was 0.93 A. A comparison of the hydrophobic energies, surface accessibility index, and the distribution of hydrophilic and hydrophobic residues of the CAD ATCase structure with correctly and incorrectly folded proteins and with several X-ray structures supported the validity of the model. The mammalian ATCase domain associates to form a compact globular trimer, a prerequisite for catalysis since the active site is comprised of residues from adjacent subunits. Interactions between the clearly defined aspartate and carbamyl phosphate subdomains of the monomer were largely preserved while there was appreciable remodeling of the trimeric interfaces. Several clusters of basic residues are located on the upper surface of the domain which account in part for the elevated isoelectric point (pI = 9.4) and may represent contact regions with other more acidic domains within the chimeric polypeptide. A long interdomain linker connects the monomer at its upper surface to the remainder of the polypeptide. The configuration of active site residues is virtually identical in the mammalian and bacterial enzymes. While the CAD ATCase domain can undergo the local conformational changes that accompany catalysis in the E. coli enzyme, the high activity, closed conformation is probably more stable in the mammalian enzyme.  相似文献   

5.
P F Lue  D M Aitken  J G Kaplan 《Biochimie》1976,58(1-2):19-25
Kinetic studies of the carbamyl phosphate synthetase activity (CPSase) of bakers' yeast revealed an absolute requirement for K+ ions ; KM values for two of the substrates, glutamine and bicarbonate, were found to be 5 X 10(-4) M and 3 X 10(-3) M respectively. CPSase activity of the purified enzyme aggregate (M.W. 800,000) was extremely sensitive to UTP with a Ki of 2.4 X 10(-4) M. The purine nucleotide intermediate, XMP, was a strong activator of CPSase, acting at a site different from the regulatory site at which UTP binds ; XMP activation diminished at high concentrations of the substrate Mg-ATP. Studies of the reaction mechanism of CPSase revealed that it involved the sequential addition of the substrates bicarbonate and Mg-ATP, liberation of ADP, addition of glutamine, binding of ATP and then release of ADP and the product carbamyl phosphate. Studies of the reaction mechanism of the aspartate transcarbamylase (ATCase) of the aggregate yielded data which were not compatible with any of the usual models ; whichever reaction mechanism is ultivately found to fit the data, it will probably prove applicable both to the ATCase of the aggregate and to the disaggregated ATCase subunit (MW 138,000).  相似文献   

6.
The pathway of product release from the R state of aspartate transcarbamoylase (ATCase; EC 2.1.3.2, aspartate carbamoyltransferase) has been determined here by solving the crystal structure of Escherichia coli ATCase locked in the R quaternary structure by specific introduction of disulfide bonds. ATCase displays ordered substrate binding and product release, remaining in the R state until substrates are exhausted. The structure reported here represents ATCase in the R state bound to the final product molecule, phosphate. This structure has been difficult to obtain previously because the enzyme relaxes back to the T state after the substrates are exhausted. Hence, cocrystallizing the wild-type enzyme with phosphate results in a T-state structure. In this structure of the enzyme trapped in the R state with specific disulfide bonds, we observe two phosphate molecules per active site. The position of the first phosphate corresponds to the position of the phosphate of carbamoyl phosphate (CP) and the position of the phosphonate of N-phosphonacetyl-l-aspartate. However, the second, more weakly bound phosphate is bound in a positively charged pocket that is more accessible to the surface than the other phosphate. The second phosphate appears to be on the path that phosphate would have to take to exit the active site. Our results suggest that phosphate dissociation and CP binding can occur simultaneously and that the dissociation of phosphate may actually promote the binding of CP for more efficient catalysis.  相似文献   

7.
Carbamyl phosphate synthase-I and glutamate dehydrogenase both form a complex with mitochondrial aspartate aminotransferase. Instead of these two enzymes competing for the aminotransferase, carbamyl phosphate synthase-I enhances glutamate dehydrogenase-aminotransferase interaction. This suggests that a complex can be formed between all three enzymes. Since this complex is stable in the presence of substrates and modifiers of the three enzymes, it could conceivably convert NH4+ produced from aspartate into carbamyl phosphate. Furthermore, since carbamyl phosphate synthase-I is the predominant protein in liver mitochondria, it could play a major role in placing the aminotransferase and glutamate dehydrogenase in close proximity. Malate removes glutamate dehydrogenase from the tri-enzyme complex and thus could play a role in determining whether glutamate dehydrogenase interacts with carbamyl phosphate synthase-I or is available to participate in reactions with the Krebs cycle. Palmitoyl-CoA has a high affinity for both carbamyl phosphate synthase-I and glutamate dehydrogenase. ATP and malate which, respectively, decrease and enhance binding of palmitoyl-CoA to glutamate dehydrogenase, respectively decrease and enhance the ability of this enzyme to compete with carbamyl phosphate synthase-I for palmitoyl-CoA. Since carbamyl phosphate synthase-I is present in high levels in liver mitochondria and has a high affinity for palmitoyl-CoA, it could play a major role as a reservoir for palmitoyl-CoA.  相似文献   

8.
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   

9.
Some metabolic pathways are nearly ubiquitous among organisms: the genes encoding the enzymes for such pathways must therefore be ancient and essential. De novo pyrimidine biosynthesis is an example of one such metabolic pathway. In animals a single protein called CAD
  • 1 Abbreviations: CAD, trifunctional protein catalyzing the first three steps of de novo pyrimidine biosynthesis in higher eukaryotes; CPS, carbamyl phosphate synthetase domain; CPSase, carbamyl phosphate synthetase activity; ATC, aspartate transcarbamylase domain; ATCase, aspartate transcarbamylase activity; DHO, dihydroorotase domain; DHOase, dihydroorotase activity; GLN, glutaminase subdomain or subunit of carbamyl phosphate synthetase, GL Nase, glutaminase activity; SYN, synthetase subdomain or subunit of carbamyl phosphate synthetase; SYNase, synthetase activity.
  • carries the first three steps of this pathway. The same three enzymes in prokaryotes are associated with separate proteins. The CAD gene appears to have evolved through a process of gene duplication and DNA rearrangement, leading to an in-frame gene fusion encoding a chimeric protein. A driving force for the creation of eukaryotic genes encoding multienzymatic proteins such as CAD may be the advantage of coordinate expression of enzymes catalyzing steps in a biosynthetic pathway. The analogous structure in bacteria is the operon. Differences in the translational mechanisms of eukaryotes and prokaryotes may have dictated the different strategies used by organisms to evolve coordinately regulated genes.  相似文献   

    10.
    Previous pKa determinations indicated that histidine 134, present in the catalytic site of aspartate transcarbamylase, might be the group involved in the binding of the substrate carbamyl phosphate and, possibly, in the catalytic efficiency of this enzyme. In the present work, this residue was replaced by an asparagine through site-directed mutagenesis. The results obtained show that histidine 134 is indeed the group of the enzyme whose deprotonation increases the affinity of the catalytic site for carbamyl phosphate. In the wild-type enzyme this group can be titrated only by those carbamyl phosphate analogues that bear the carbonyl group. In the modified enzyme the group whose deprotonation increases the catalytic efficiency is still present, indicating that this group is not the imidazole ring of histidine 134 (pKa = 6.3). In addition, the pKa of the still unknown group involved in aspartate binding is shifted by one unit in the mutant as compared to the wild type.  相似文献   

    11.
    Site-directed mutagenesis was used to create four mutant versions of Escherichia coli aspartate transcarbamylase at three positions in the catalytic chain of the enzyme. The location of all the amino acid substitutions was near the carbamyl phosphate binding site as previously determined by X-ray crystallography. Arg-54, which interacts with both the anhydride oxygen and a phosphate oxygen of carbamyl phosphate, was replaced by alanine. This mutant enzyme was approximately 17,000-fold less active than the wild type, although the binding of substrates and substrate analogues was not altered substantially. Arg-105, which interacts with both the carbonyl oxygen and a phosphate oxygen of carbamyl phosphate, was replaced by alanine. This mutant enzyme exhibited an approximate 1000-fold loss of activity, while the activity of catalytic subunit isolated from this mutant enzyme was reduced by 170-fold compared to the wild-type catalytic subunit. The KD of carbamyl phosphate and the inhibition constants for acetyl phosphate and N-(phosphono-acetyl)-L-aspartate (PALA) were increased substantially by this amino acid substitution. Furthermore, this loss in substrate and substrate analogue binding can be correlated with the large increases in the aspartate and carbamyl phosphate concentrations at half of the maximum observed specific activity, [S]0.5. Gln-137, which interacts with the amino group of carbamyl phosphate, was replaced by both asparagine and alanine. The asparagine mutant exhibited only a small reduction in activity while the alanine mutant was approximately 50-fold less active than the wild type. The catalytic subunits of both these mutant enzymes were substantially more active than the corresponding holoenzymes.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

    12.
    The optimal conditions for storing and assaying mesozooplanktonaspartate transcarb amylase (ATCase) (EC 2.1.3.1 [EC] .) are definedin order to enable the specific activity of this enzyme to beused as an index of mesozooplankton productivity. ATCase activitywas found to be stable for more than 2 weeks when stored inliquid nitrogen, with a negligible loss of activity, whateverthe state of frozen material (cell-free homogenate or wholecells). As an alternative, –20C, –90C freezersand dry ice (–80C) can give more flexibility for storingand transporting material, provided initial freezing has beencarried out at –196C. ATCase specific activity was stableover a large range of total protein concentrations (2–55mg ml–1) for most mesozooplankton samples, and the enzymeactivity was linear between 0.8 and 80 nmol carbamyl aspartatemin–1 mg–1 of total protein. The optimal incubationtime (30 min), temperature (35C) and pH range (9–9.5)for assay of ATCase were similar forCalanus helgolondicus (Crustacea:Copepoda) (G.O.Sars) and three mesozooplankton assemblages sampledoff Plymouth (UK) in May, August and November. However, theKm values for both substrateswere variable: from 3.9 to 79.1for aspartate, and from 0.83 to 3.46 for carbamyl phosphate.Such variations arelikely to be due to the presence of differentATCase catalytic potential, rather than changes in ATCase assemblage.  相似文献   

    13.
    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   

    14.
    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   

    15.
    Allosteric enzyme models on the basis of the known properties of aspartate transcarbamylase (ATCase) from Escherichia coli are suggested. In the first model molecules are supposed to equilibrate between two states. In contrast to the classical Monod-Wyman-Changeux model the symmetry of enzyme molecules changes during the conformational transition. It is shown that the number of binding sites of the enzyme defined from the Scatchard plots is sufficiently dependent on values of parameters of enzyme reaction. This fact results from the mixed (both positive and negative) cooperative effects. However the complex kinetic of ATCase is not completely simulated by this model. Therefore the model is complicated by taking into account the inactive third state of the enzyme. Thus the complex kinetic behaviour of ATCase is explained. The models may be also used for other enzymes.  相似文献   

    16.
    We have studied the kinetics of the quaternary structure change associated with the allosteric transition of aspartate transcarbamylase (ATCase) (E. coli), inducing this change by exposure to the natural substrates (carbamyl phosphate and L-aspartate). The presence of 30% ethylene glycol slowed the quaternary structure change sufficiently for it to be followed by stopped-flow X-ray scattering at -5 degrees C. After adding substrates to the enzyme, the change occurred, with a half-life of a few seconds, yielding a mixture of the two standard quaternary structures (or, conceivably, a state intermediate between them). This mixture persisted until the enzyme reduced the substrate concentration below a threshold value.  相似文献   

    17.
    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   

    18.
    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   

    19.
    Aquifex aeolicus, an organism that flourishes at 95 degrees C, is one of the most thermophilic eubacteria thus far described. The A. aeolicus pyrB gene encoding aspartate transcarbamoylase (ATCase) was cloned, overexpressed in Escherichia coli, and purified by affinity chromatography to a homogeneous form that could be crystallized. Chemical cross-linking and size exclusion chromatography showed that the protein was a homotrimer of 34-kDa catalytic chains. The activity of A. aeolicus ATCase increased dramatically with increasing temperature due to an increase in kcat with little change in the Km for the substrates, carbamoyl phosphate and aspartate. The Km for both substrates was 30-40-fold lower than the corresponding values for the homologous E. coli ATCase catalytic subunit. Although rapidly degraded at high temperature, the carbamoyl phosphate generated in situ by A. aeolicus carbamoyl phosphate synthetase (CPSase) was channeled to ATCase. The transient time for carbamoyl aspartate formation was 26 s, compared with the much longer transient times observed when A. aeolicus CPSase was coupled to E. coli ATCase. Several other approaches provided strong evidence for channeling and transient complex formation between A. aeolicus ATCase and CPSase. The high affinity for substrates combined with channeling ensures the efficient transfer of carbamoyl phosphate from the active site of CPSase to that of ATCase, thus preserving it from degradation and preventing the formation of toxic cyanate.  相似文献   

    20.
    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   

    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