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 368 毫秒
1.
The prephenate dehydrogenase activity of the bifunctional enzyme chorismate mutase-prephenate dehydrogenase from Escherichia coli catalyzes the oxidative decarboxylation of both prephenate and deoxoprephenate, which lacks the keto group in the side chain (V 78% and V/K 18% those of prephenate). Hydride transfer is to the B side of NAD, and the acetylpyridine and pyridinecarboxaldehyde analogues of NAD have V/K values 40 and 9% and V values 107 and 13% those of NAD. Since the 13C isotope effect on the decarboxylation is 1.0103 with deuterated and 1.0033 with unlabeled deoxoprephenate (the deuterium isotope effect on V/K is 2.34), the mechanism is concerted, and if CO2 has no reverse commitment, the intrinsic 13C and deuterium isotope effects are 1.0155 (corresponding to a very early transition state for C-C bond cleavage) and 7.3, and the forward commitment is 3.7. With deoxodihydroprephenate (lacking one double bond in the ring), oxidation occurs without decarboxylation, and one enantiomer has a V/K value 23-fold higher than the other (deuterium isotope effects are 3.6 and 4.1 for fast and slow isomers; V for the fast isomer is 5% and V/K 0.7% those of prephenate). The fully saturated analogue of deoxoprephenate is a very slow substrate (V 0.07% and V/K approximately 10(-5%) those of prephenate). pH profiles show a group with pK = 8.3 that must be protonated for substrate binding and a catalytic group with pK = 6.5 that is a cationic acid (likely histidine). This group facilitates hydride transfer by beginning to accept the proton from the 4-hydroxyl group of prephenate prior to the beginning of C-C cleavage (or fully accepting it in the oxidation of the analogues with only one double bond or none in the ring). In contrast with the enzymatic reaction, the acid-catalyzed decarboxylation of prephenate and deoxoprephenate (t1/2 of 3.7 min at low pH) is a stepwise reaction with a carbonium ion intermediate, since 18O is incorporated into substrate and its epi isomer during reaction in H218O. pH profiles show that the hydroxyl group must be protonated and the carboxyl (pK approximately 4.2) ionized for carbonium ion formation. The carbonium ion formed from prephenate decarboxylates 1.75 times faster than it reacts with water (giving 1.8 times as much prephenate as epi isomer). The observed 13C isotope effect of 1.0082 thus corresponds to an intrinsic isotope effect of 1.023, indicating an early transition state for the decarboxylation step. epi-Prephenate is at least 20 times more stable to acid than prephenate because it exists largely as an internal hemiketal.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

2.
The variation with pH of the kinetic parameters associated with the mutase and dehydrogenase reactions catalyzed by chorismate mutase-prephenate dehydrogenase has been determined with the aim of elucidating the role that ionizing amino acid residues play in binding and catalysis. The pH dependency of log V for the dehydrogenase reaction shows that the enzyme possesses a single ionizing group with a pK value of 6.5 that must be unprotonated for catalysis. This same group is observed in the V/Kprephenate, as well as in the V/KNAD, profile. The V/Kprephenate profile exhibits a second ionizing residue with a pK value of 8.4 that must be protonated for the binding of prephenate to the enzyme. For the mutase reaction, the V/Kchorismate profile indicates the presence of three ionizing residues at the active site. Two of these residues, with similar pK values of about 7, must be protonated, while the third, with a pK value of 6.3, must be unprotonated. It can be concluded that all three groups are concerned with the binding of chorismate to the enzyme since the maximum velocity of the mutase reaction is essentially independent of pH. This conclusion is confirmed by the finding that the Ki profile for the competitive inhibitor, (3-endo,8-exo)-8-hydroxy-2-oxabicyclo[3.3]non-6-ene-3,5-dicarboxylic acid, shows the same three ionizing groups as observed in the V/Kchorismate profile. By contrast, the Ki profile for carboxyethyldihydrobenzoate shows only one residue, with a pK value of 7.3, that must be protonated for binding of the inhibitor.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
Three classes of mutant strains of Escherichia coli K12 defective in pheA, the gene coding for chorismate mutase/prephenate dehydratase, have been isolated: (1) those lacking prephenate dehydratase activity, (2) those lacking chorismate mutase activity, and (3) those lacking both activities. Chorismate mutase/prephenate dehydratase from the second class of mutants was less sensitive to inhibition by phenylalanine than wild-type enzyme and, along with the defective enzyme from the third class of mutants, could not be purified by affinity chromatography on Sepharosyl-phenylalanine. Pure chorismate mutase/prephenate dehydratase protein was prepared from two strains belonging to the first class. The chorismate mutase activity of these enzymes is kinetically similar to that of the wild-type enzyme except for a two- to threefold increase in both the Ka for chorismate and the Kis for inhibition by prephenate. In both cases only one change in the tryptic fingerprint was detected, resulting from a substitution of the threonine residue in the peptide Gln·Asn·Phe·Thr·Arg. This suggests that this residue is catalytically or structurally essential for the dehydratase activity.  相似文献   

4.
The enzyme activities specified by the tyrA and pheA genes were studied in wildtype strain Salmonella typhimurium and in phenylalanine and tyrosine auxotrophs. As in Aerobacter aerogenes and Escherichia coli, the wild-type enzymes of Salmonella catalyze two consecutive reactions: chorismate --> prephenate --> 4-hydroxy-phenylpyruvate (tyrA), and chorismate --> prephenate --> phenylpyruvate (pheA). A group of tyrA mutants capable of interallelic complementation had altered enzymes which retained chorismate mutase T activity but lacked prephenate dehydrogenase. Similarly, pheA mutants (in which interallelic complementation does not occur) had one group with altered enzymes which retained chorismate mutase P but lacked prephenate dehydratase. Tyrosine and phenylalanine auxotrophs outside of these categories showed loss of both activities of their respective bifunctional enzyme. TyrA mutants which had mutase T were considerably derepressed in this activity by tyrosine starvation and consequently excreted prephenate. A new and specific procedure was developed for assaying prephenate dehydrogenase activity.  相似文献   

5.
Pig heart mitochondrial NADP-dependent isocitrate dehydrogenase is the most extensively studied among the mammalian isocitrate dehydrogenases. The crystal structure of Escherichia coli isocitrate dehydrogenase and sequence alignment of porcine with E. coli isocitrate dehydrogenase suggests that the porcine Arg(101), Arg(110), Arg(120), and Arg(133) are candidates for roles in substrate binding. The four arginines were separately mutated to glutamine using a polymerase chain reaction method. Wild type and mutant enzymes were each expressed in E. coli, isolated as maltose binding fusion proteins, then cleaved with thrombin, and purified to yield homogeneous porcine isocitrate dehydrogenase. The R120Q mutant has a specific activity, as well as K(m) values for isocitrate, Mn(2+), and NADP(+) similar to wild type enzyme, indicating that Arg(120) is not needed for function. The specific activities of R101Q, R110Q, and R133Q are 1.73, 1.30, and 19.7 micromols/min/mg, respectively, as compared with 39.6 units/mg for wild type enzyme. The R110Q and R133Q enzymes exhibit K(m) values for isocitrate that are increased more than 400- and 165-fold, respectively, as compared with wild type. The K(m) values for Mn(2+), but not for NADP(+), are also elevated indicating that binding of the metal-isocitrate complex is impaired in these mutants. It is proposed that the positive charges of Arg(110) and Arg(133) normally strengthen the binding of the negatively charged isocitrate by electrostatic attraction. The R101Q mutant shows smaller, but significant increases in the K(m) values for isocitrate and Mn(2+); however, the marked decrease in k(cat) suggests a role for Arg(101) in catalysis. The V(max) of wild type enzyme depends on the ionized form of an enzymic group of pK 5.5, and this pK(aes) is similar for the R101Q and R120Q enzymes. In contrast, the pK(aes) for R110Q and R133Q enzymes increases to 6.4 and 7.4, respectively, indicating that the positive charges of Arg(110) and Arg(133) normally lower the pK of the nearby catalytic base to facilitate its ionization. These results may be understood in terms of the structure of the porcine NADP-specific isocitrate dehydrogenase generated by the Insight II Modeler Program, based on the x-ray coordinates of the E. coli enzyme.  相似文献   

6.
The enzyme prephenate dehydrogenase catalyzes the oxidative decarboxylation of prephenate to 4-hydroxyphenylpyruvate for the biosynthesis of tyrosine. Prephenate dehydrogenases exist as either monofunctional or bifunctional enzymes. The bifunctional enzymes are diverse, since the prephenate dehydrogenase domain is associated with other enzymes, such as chorismate mutase and 3-phosphoskimate 1-carboxyvinyltransferase. We report the first crystal structure of a monofunctional prephenate dehydrogenase enzyme from the hyper-thermophile Aquifex aeolicus in complex with NAD+. This protein consists of two structural domains, a modified nucleotide-binding domain and a novel helical prephenate binding domain. The active site of prephenate dehydrogenase is formed at the domain interface and is shared between the subunits of the dimer. We infer from the structure that access to the active site is regulated via a gated mechanism, which is modulated by an ionic network involving a conserved arginine, Arg250. In addition, the crystal structure reveals for the first time the positions of a number of key catalytic residues and the identity of other active site residues that may participate in the reaction mechanism; these residues include Ser126 and Lys246 and the catalytic histidine, His147. Analysis of the structure further reveals that two secondary structure elements, beta3 and beta7, are missing in the prephenate dehydrogenase domain of the bifunctional chorismate mutase-prephenate dehydrogenase enzymes. This observation suggests that the two functional domains of chorismate mutase-prephenate dehydrogenase are interdependent and explains why these domains cannot be separated.  相似文献   

7.
Karsten WE  Liu D  Rao GS  Harris BG  Cook PF 《Biochemistry》2005,44(9):3626-3635
The pH dependence of kinetic parameters of several active site mutants of the Ascaris suum NAD-malic enzyme was investigated to determine the role of amino acid residues likely involved in catalysis on the basis of three-dimensional structures of malic enzyme. Lysine 199 is positioned to act as the general base that accepts a proton from the 2-hydroxyl of malate during the hydride transfer step. The pH dependence of V/K(malate) for the K199R mutant enzyme reveals a pK of 5.3 for an enzymatic group required to be unprotonated for activity and a second pK of 6.3 that leads to a 10-fold loss in activity above the pK of 6.3 to a new constant value up to pH 10. The V profile for K199R is pH independent from pH 5.5 to pH 10 and decreases below a pK of 4.9. Tyrosine 126 is positioned to act as the general acid that donates a proton to the enolpyruvate intermediate to form pyruvate. The pH dependence of V/K(malate) for the Y126F mutant is qualitatively similar to K199R, with a requirement for a group to be unprotonated for activity with a pK of 5.6 and a partial activity loss of about 3-fold above a pK of 6.7 to a new constant value. The Y126F mutant enzyme is about 60000-fold less active than the wild-type enzyme. In contrast to K199R, the V rate profile for Y126F also shows a partial activity loss above pH 6.6. The wild-type pH profiles were reinvestigated in light of the discovery of the partial activity change for the mutant enzymes. The wild-type V/K(malate) pH-rate profile exhibits the requirement for a group to be unprotonated for catalysis with a pK of 5.6 and also shows the partial activity loss above a pK of 6.4. The wild-type V pH-rate profile decreases below a pK of 5.2 and is pH independent from pH 5.5 to pH 10. Aspartate 294 is within hydrogen-bonding distance to K199 in the open and closed forms of malic enzyme. D294A is about 13000-fold less active than the wild-type enzyme, and the pH-rate profile for V/K(malate) indicates the mutant is only active above pH 9. The data suggest that the pK present at about pH 5.6 in all of the pH profiles represents D294, and during catalysis D294 accepts a proton from K199 to allow K199 to act as a general base in the reaction. The pK for the general acid in the reaction is not observed, consistent with rapid tautomerization of enolpyruvate. No other ionizable group in the active site is likely responsible for the partial activity change observed in the pH profiles, and thus the group responsible is probably remote from the active site and the effect on activity is transmitted through the protein by a conformational change.  相似文献   

8.
Previous alanine scanning mutagenesis of ADP-glucose pyrophosphorylase from Anabaena PCC 7120 indicated that Arg(294) plays a role in inhibition by orthophosphate [J. Sheng, J. Preiss, Biochemistry 36 (1997) 13077]. In this study, analysis of several site-directed mutants in the presence of different metabolic effectors showed that the primary inhibitor for two of the mutant proteins, R294A and R294Q, was no longer orthophosphate but rather NADPH, which was a reversal in the pattern of inhibitor selectivity from the wild-type. Despite the differences in charge and size, analysis of the purified R294K, R294E, and R294Q mutant enzymes demonstrated similar decreases in orthophosphate affinity as the R294A mutant, while most of the other kinetic values were similar to those reported for the wild-type. All these results suggest that the positive charge of Arg(294) is not specifically involved in orthophosphate binding and that it is important in determining inhibitor selectivity.  相似文献   

9.
The regulatory properties of three key enzymes in the phenylalanine biosynthetic pathway, 3-deoxy-D-arabino-heptulosonate 7-phosphate synthetase (DAHP synthetase) [EC 4.1.2.15], chorismate mutase [EC 5.4.99.5], and prephenate dehydratase [prephenate hydro-lyase (decarboxylating), EC 4.2.1.51] were compared in three phenylalanine-excreting mutants and the wild strain of Brevibacterium flavum. Regulation of DAHP synthetase by phenylalanine and tyrosine in these mutants did not change at all, but the specific activities of the mutant cell extracts increased 1.3- to 2.8-fold, as reported previously (1). Chorismate mutase activities in both the wild and the mutant strains were cumulatively inhibited by phenylalanine and tyrosine and recovered with tryptophan, while the specific activities of the mutants increased 1.3- to 2.8-fold, like those of DAHP synthetase. On the other hand, the specific activities of prephenate dehydratase in the mutant and wild strains were similar, when tyrosine was present. While prephenate dehydratase of the wild strain was inhibited by phenylalanine, tryptophan, and several phenylalanine analogues, the mutant enzymes were not inhibited at all but were activated by these effectors. Tyrosine activated the mutant enzymes much more strongly than the wild-type enzyme: in mutant 221-43, 1 mM tyrosine caused 28-fold activation. Km and the activation constant for tyrosine were slightly altered to a half and 6-fold compared with the wild-type enzyme, respectively, while the activation constants for phenylalanine and tryptophan were 500-fold higher than the respective inhibition constants of the wild-type enzyme. The molecular weight of the mutant enzyme was estimated to be 1.2 x 10(5), a half of that of the wild-type enzyme. The molecular weight of the mutant enzyme was estimated to be 1.2 X 10(5) a half of that of the wild type enzyme, while in the presence of tyrosine, phenylalanine, or tryptophan, it increased to that of the wild-type enzyme. Immediately after the mutant enzyme had been activated by tyrosine and then the tyrosine removed, it still showed about 10-fold higher specific activity than before the activation by tyrosine. However, on standing in ice the activity gradually fell to the initial level before the activation by tyrosine. Ammonium sulfate promoted the decrease of the activity. On the basis of these results, regulatory mechanisms for phenylalanine biosynthesis in vivo as well as mechanisms for the phenylalanine overproduction in the mutants are discussed.  相似文献   

10.
S Ahmad  R A Jensen 《FEBS letters》1987,216(1):133-139
The prephenate dehydrogenase component of the bifunctional T-protein (chorismate mutase:prephenate dehydrogenase) has been shown to utilize L-arogenate, a common precursor of phenylalanine and tyrosine in nature, as a substrate. Partially purified T-protein from Klebsiella pneumoniae and from Escherichia coli strains K 12, B, C and W was used to demonstrate the utilization of L-arogenate as an alternative substrate for prephenate in the presence of nicotinamide adenine dinucleotide as cofactor. The formation of L-tyrosine from L-arogenate by the T-protein dehydrogenase was confirmed by high-performance liquid chromatography. As expected of a common catalytic site, dehydrogenase activity with either prephenate or L-arogenate was highly sensitive to inhibition by L-tyrosine.  相似文献   

11.
Adenylosuccinate lyase (ASL) of Bacillus subtilis contains three conserved histidines, His(68), His(89), and His(141), identified by affinity labeling and site-directed mutagenesis as critical to the intersubunit catalytic site. The pH-V(max) profile for wild-type ASL is bell-shaped (pK (1) = 6.74 and pK (2) = 8.28). Only the alkaline side changes with temperature, characteristic of histidine pKs. To identify determinants of pK (2) in the enzyme-substrate complex, we replaced residues at two positions close to His(68) (but not to His(89) or His(141)) in the structure. Compared with the specific activity of 1.75 mumol adenylosuccinate reacting/min/mg of wild-type enzyme at pH 7.0, mutant enzymes D69E, D69N, R310Q, and R310K exhibit specific activities of 0.40, 0.04, 0.00083, and 0.10, respectively. While D69E has a K (m) for adenylosuccinate similar to that of wild-type ASL, D69N and R310K exhibit modest increases in K (m), and R310Q has an 11-fold increase in K (m). The mutant enzymes show no significant change in molecular weight or secondary structure. The major change is in the pH-V(max) profile: pK (2) is 8.48 for the D69E mutant and is decreased to 7.83 in D69N, suggesting a proximal negative charge is needed to maintain the high pK of 8.28 observed for wild-type enzyme and attributed to His(68). Similarly, R310Q exhibits a decrease in its pK (2) (7.33), whereas R310K shows little change in pK (2) (8.24). These results suggest that Asp(69) interacts with His(68), that Arg(310) interacts with and orients the beta-carboxylate of Asp(69), and that His(68) must be protonated for ASL to be active.  相似文献   

12.
Segall ML  Colman RF 《Biochemistry》2004,43(23):7391-7402
In adenylosuccinate lyase from Bacillus subtilis, Gln(212), Asn(270), and Arg(301) are conserved and located close to the succinyl moiety of docked adenylosuccinate. We constructed mutant enzymes with Gln(212) replaced by Glu and Met, Asn(270) by Asp and Leu, and Arg(301) by Gln or Lys. The wild-type and mutant enzymes were expressed in Escherichia coli and purified to homogeneity. The specific activities of the Q212M and the 270 and 301 mutant enzymes were decreased more than 3000-fold as compared to the wild type. Only Q212E retained sufficient activity for determination of its kinetic parameters: V(max) was decreased approximately 1000-fold, and K(m) was increased 6-fold, as compared to the wild-type enzyme. Adenylosuccinate binding studies of the other mutants revealed greatly weakened affinities that contributed to, but did not account entirely for, the loss of activity. These mutant enzymes did not differ greatly from the wild-type enzyme in secondary structure or subunit association state, as shown by circular dichroism spectroscopy and light-scattering photometry. Incubation of pairs of inactive mutant enzymes led to reconstitution of some functional sites by subunit complementation, with recovery of up to 25% of the specific activity of the wild-type enzyme. Subunit complementation occurs only if the two mutations are contributed to the active site by different subunits. Thus, mixing Q212E with N270L enzyme yielded a specific activity of approximately 20% of the wild-type enzyme, while mixing Q212M with R301K enzyme did not restore activity. As supported by computer modeling, the studies presented here indicate that Gln(212), Asn(270), and Arg(301) are indispensable to catalysis by adenylosuccinate lyase and probably interact noncovalently with the carboxylate anions of the substrates 5-aminoimidazole-4(N-succinylocarboxamide)ribonucleotide and adenylosuccinate, optimizing their bound orientations.  相似文献   

13.
Xia L  Ballou DP  Marsh EN 《Biochemistry》2004,43(11):3238-3245
Arginine-100 is involved in recognizing the gamma carboxylate of the substrate in glutamate mutase. To investigate its role in substrate binding and catalysis, this residue was mutated to lysine, tyrosine, and methionine. The effect of these mutations was to reduce k(cat) by 120-320-fold and to increase K(m(apparent)) for glutamate by 13-22-fold; K(m(apparent)) for adenosylcobalamin is little changed by these mutations. Even at saturating substrate concentrations, no cob(II)alamin could be detected in the UV-visible spectra of the Arg100Tyr and Arg100Met mutants. However, in the Arg100Lys mutant cob(II)alamin accumulated to concentrations similar to wild-type enzyme, which allowed the pre-steady-state kinetics of adenosylcobalamin homolysis to be investigated by stopped-flow spectroscopy. It was found that homolysis of the coenzyme is slower by an order of magnitude, compared with wild-type enzyme. Furthermore, glutamate binding is significantly weakened, so much so that the reaction exhibits second-order kinetics over the range of substrate concentrations used. The Arg100Lys mutant does not exhibit the very large deuterium isotope effects that are observed for homolysis of the coenzyme when the wild-type enzyme is reacted with deuterated substrates; this suggests that homolysis is slowed relative to hydrogen abstraction by this mutation.  相似文献   

14.
The effect of pH on chorismate mutase/prephenate dehydratase (chorismate pyruvate mutase/prephenate hydro-lyase (decarboxylating) EC 5.4.99.5/EC 4.2.1.51) from Escherichia coli K12 has been studied. While the maximum velocity of both activities is independent of pH, Km for chorismate or prephenate shows a complex pH dependence. Differences in mutase activity in acetate/phosphate/borate and citrate/phosphate/borate buffers were traced to inhibition by citrate. When a variety of analogues of citrate were tested as possible inhibitors of the enzyme, several were found to inhibit mutase and dehydratase activities to different extents, and by different mechanisms. Thus citrate competitively inhibits mutase activity, but inhibits dehydratase activity by either a non-competitive or an uncompetitive mechanism. Conversely, cis- and trans-aconitate competitively inhibit dehydratase activity, but are partially competitive inhibitors of mutase activity. The differential effects of these inhibitors on the two activities are consistent with the existence of two distinct active sites, but additionally suggest some degree of interconnection between them. The implications of these results for possible mechanisms of catalysis by chorismate mutase/prephenate dehydratase are discussed.  相似文献   

15.
The chlorite dismutase from Dechloromonas aromatica (DaCld) catalyzes the highly efficient decomposition of chlorite to O(2) and chloride. Spectroscopic, equilibrium thermodynamic, and kinetic measurements have indicated that Cld has two pH sensitive moieties; one is the heme, and Arg183 in the distal heme pocket has been hypothesized to be the second. This active site residue has been examined by site-directed mutagenesis to understand the roles of positive charge and hydrogen bonding in O-O bond formation. Three Cld mutants, Arg183 to Lys (R183K), Arg183 to Gln (R183Q), and Arg183 to Ala (R183A), were investigated to determine their respective contributions to the decomposition of chlorite ion, the spin state and coordination states of their ferric and ferrous forms, their cyanide and imidazole binding affinities, and their reduction potentials. UV-visible and resonance Raman spectroscopies showed that DaCld(R183A) contains five-coordinate high-spin (5cHS) heme, the DaCld(R183Q) heme is a mixture of five-coordinate and six-coordinate high spin (5c/6cHS) heme, and DaCld(R183K) contains six-coordinate low-spin (6cLS) heme. In contrast to wild-type (WT) Cld, which exhibits pK(a) values of 6.5 and 8.7, all three ferric mutants exhibited pH-independent spectroscopic signatures and kinetic behaviors. Steady state kinetic parameters of the chlorite decomposition reaction catalyzed by the mutants suggest that in WT DaCld the pK(a) of 6.5 corresponds to a change in the availability of positive charge from the guanidinium group of Arg183 to the heme site. This could be due to either direct acid-base chemistry at the Arg183 side chain or a flexible Arg183 side chain that can access various orientations. Current evidence is most consistent with a conformational adjustment of Arg183. A properly oriented Arg183 is critical for the stabilization of anions in the distal pocket and for efficient catalysis.  相似文献   

16.
Zhang S  Wilson DB  Ganem B 《Biochemistry》2000,39(16):4722-4728
The Escherichia coli bifunctional P-protein, which plays a central role in L-phenylalanine (Phe) biosynthesis, contains distinct chorismate mutase (CM) and prephenate dehydratase (PDT) domains as well as a regulatory (R) domain for feedback control by Phe. To elucidate the catalytic mechanism of PDT in the P-protein, 24 mutations of 15 conserved residues in the PDT domain were created, expressed in the pheA(-)E. coli strain NK6024, and studied for their effect on PDT activity. Fourteen mutant enzymes were purified to homogeneity, tested for feedback inhibition by Phe, and characterized by kinetic analysis and circular dichroism spectroscopy. Selected mutant enzymes were further studied by gel filtration, fluorescence emission, and microcalorimetry. In addition, a monofunctional PDT domain (PDT20, residues 101-285) was cloned and overexpressed in plasmid pET with expression levels up to 200-250 mg/L. PDT20 retained full PDT activity, lacked CM activity, and was insensitive to feedback inhibition by Phe. Four residues (T278, N160, Q215, and S208) were shown to be important for PDT catalysis. The values of k(cat)/K(m) for the S208A/C and T278S mutant enzymes were 100-fold lower, and 500-fold lower for the N160A and Q215A mutant enzymes than the wild-type (WT) protein. The T278A and T278V mutant enzymes displayed no measurable catalytic activity, yet bound both prephenate and a competitive inhibitor (S-DNBA) comparably to the WT protein. These data, taken together with the normal CD spectra of the mutant enzymes, strongly suggested that T278 was involved in the catalytic mechanism. To establish whether acidic residues were involved in catalysis, all the conserved Glu and Asp residues in the PDT domain were mutated to Ala. None of these mutations significantly reduced PDT activity, indicating that the acidic residues of the PDT domain are not directly involved in catalysis. However, two mutant enzymes (E159A and E232A) displayed higher levels of PDT activity (2.2- and 3.5-fold, respectively), which was due to enhanced substrate binding. For the double mutant enzyme (E159A-E232A), k(cat)/K(m) was ca. 7-fold higher than for the WT enzyme, while its K(m) was 4.6-fold lower.  相似文献   

17.
Aminopeptidases are major enzymes in the midgut microvillar membranes of most insects and are targets of insecticidal Bacillus thuringiensis crystal delta-endotoxins. Sequence analysis and substrate specificity studies showed that these enzymes resemble mammalian aminopeptidase N, although information on the organization of their active site is lacking. The effect of pH at different temperatures on the kinetic parameters of Tenebrio molitor (Coleoptera) larval aminopeptidase showed that enzyme catalysis depend on a deprotonated (pK 7.6; DeltaH degrees (ion), 7.6 kJ/mol) and a protonated (pK 8.2; DeltaH degrees (ion), 16.8 kJ/mol) group. 1-Ethyl-3-(3-dimethylaminopropyl) carbodiimide and diethylpyrocarbonate inactivate the enzyme by modifying a pK 5.8 carboxylate and a imidazole group, respectively, with a reaction order around 1. Tetranitromethane changes the K(m) of the enzyme without affecting its V(max) by modifying a phenol group. The presence of a competitive inhibitor decrease the inactivation reaction rates in all these cases. EDTA inactivation of the aminopeptidase is affected by pH and temperature suggesting the involvement in metal binding of at least one deprotonated imidazole group (pK 5.8, DeltaH degrees (ion), 20 kJ/mol). The data support the hypothesis that T. molitor aminopeptidase catalysis depends on a catalytic metal and on a carboxylate and a protonated imidazole group, whereas substrate binding relies in one phenol and one carboxylate groups. The insect aminopeptidase shares common features with mammalian aminopeptidase N, although differing in details of substrate binding and in residues directly involved in catalysis.  相似文献   

18.
R Schinzel  D Palm  K D Schnackerz 《Biochemistry》1992,31(16):4128-4133
Changes in the active site of Escherichia coli maltodextrin phosphorylase created by substituting residues Lys533, Arg534, Tyr538, and Glu637 were monitored in the absence and presence of arsenate as substrate analogue using pyridoxal-P as 31P NMR reporter. The chemical shift of the cofactor phosphate group of wild-type E. coli phosphorylase is pH dependent with an apparent pK of 5.6 and limiting delta values of 0.71 and 3.6 ppm for the low- and high-pH values, respectively. The apparent pK value of 5.6 indicates that the phosphate group of the cofactor is in hydrogen bond linkage to Lys533. In all mutant enzymes in which the enzymatic activity was significantly reduced, effects on the 31P chemical shift pattern of pyridoxal-P were observed. The K533S, R534Q, E637D, and E637Q mutant enzymes show 0.6, 0.01, 0.2, or 0.1% residual activity, and the apparent pK values of the cofactor phosphate transition of E637D and E637Q mutant enzymes are altered. The Y538F mutant enzyme is a remarkable exception, displaying 12% activity and an environment of the cofactor quite similar to that in wild-type enzyme. This finding suggests that Tyr538, although involved in substrate binding and specificity, is not functionally essential. One crucial aspect of catalysis is the close contact of the phosphates of pyridoxal-P and of substrate rendered by a cluster of positively charged amino acids, Lys533, Lys539, and Arg534. The similar apparent pK values of wild-type and K533S mutant phosphorylase suggest that the cofactor phosphate and the hydroxyl group of Ser533 are linked by a hydrogen bond.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
Highly purified enzymes from Alcaligenes eutrophus H 16 were used for kinetic studies. Chorismate mutase was feedback inhibited by phenylalanine. In the absence of the inhibitor, the double-reciprocal plot was linear, yielding a Km for chorismate of 0.2 mM. When phenylalanine was present, a pronounced deviation from the Michaelis-Menten hyperbola occurred. The Hill coefficient (n) was 1.7, and Hill plots of velocity versus inhibitor concentrations resulted in a value of n' = 2.3, indicating positive cooperativity. Chorismate mutase was also inhibited by prephenate, which caused downward double-reciprocal plots and a Hill coefficient of n = 0.7, evidence for negative cooperativity. The pH optimum of chorismate mutase ranged from 7.8 to 8.2; its temperature optimum was 47 C. Prephenate dehydratase was competitively inhibited by phenylalanine and activated by tyrosine. Tyrosine stimulated its activity up to 10-fold and decreased the Km for prephenate, which was 0.67 mM without effectors. Tryptophan inhibited the enzyme competitively. Its inhibition constant (Ki = 23 muM) was almost 10-fold higher than that determined for phenylalanine (Ki = 2.6 muM). The pH optimum of prephenate dehydratase was pH 5.7; the temperature optimum was 48 C. Prephenate dehydrogenase was feedback inhibited by tyrosine. Inhibition was competitive with prephenate (Ki = 0.06 mM) and noncompetitive with nicotinamide adenine dinucleotide. The enzyme was further subject to product inhibition by p-hydroxyphenylpyruvate (Ki = 0.13 mM). Its Km for prephenate was 0.045 mM, and that for nicotinamide adenine dinucleotide was 0.14 mM. The pH optimum ranged between 7.0 and 7.6; the temperature optimum was 38 C. It is shown how the sensitive regulation of the entire enzyme system leads to a well-balanced amino acid production.  相似文献   

20.
The inhibition of the bifunctional enzyme chorismate mutase-prephenate dehydrogenase (4-hydroxyphenylpyruvate synthase) by substrate analogues has been investigated at pH 6.0 with the aim of elucidating the spatial relationship that exists between the sites at which each reaction occurs. Several chorismate and adamantane derivatives, as well as 2-hydroxyphenyl acetate and diethyl malonate, act as linear competitive inhibitors with respect to chorismate in the mutase reaction and with respect to chorismate in the mutase reaction and with respect to prephenate in the dehydrogenase reaction. The similarity of the dissociation constants for the interaction of these compounds with the free enzyme, as determined from the mutase and dehydrogenase reactions, indicates that the reaction of these inhibitors at a single site prevents the binding of both chorismate and prephenate. However, not all the groups on the enzyme, which are responsible for the binding of these two substrates, can be identical. At lower concentrations, citrate or malonate prevents reaction of the enzyme with prephenate, but not with chorismate. Nevertheless, the combining sites for chorismate and prephenate are in such close proximity that the diethyl derivative of malonate prevents the binding of both substrates. The results lead to the proposal that the sites at which chorismate and prephenate react on hydroxyphenylpyruvate synthase share common features and can be considered to overlap.  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