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 46 毫秒
1.
Undecaprenyl pyrophosphate synthase (UPPs) catalyzes the consecutive condensation reactions of a farnesyl pyrophosphate (FPP) with eight isopentenyl pyrophosphates (IPP), in which new cis-double bonds are formed, to generate undecaprenyl pyrophosphate that serves as a lipid carrier for peptidoglycan synthesis of bacterial cell wall. The structures of Escherichia coli UPPs were determined previously in an orthorhombic crystal form as an apoenzyme, in complex with Mg(2+)/sulfate/Triton, and with bound FPP. In a further search of its catalytic mechanism, the wild-type UPPs and the D26A mutant are crystallized in a new trigonal unit cell with Mg(2+)/IPP/farnesyl thiopyrophosphate (an FPP analogue) bound to the active site. In the wild-type enzyme, Mg(2+) is coordinated by the pyrophosphate of farnesyl thiopyrophosphate, the carboxylate of Asp(26), and three water molecules. In the mutant enzyme, it is bound to the pyrophosphate of IPP. The [Mg(2+)] dependence of the catalytic rate by UPPs shows that the activity is maximal at [Mg(2+)] = 1 mm but drops significantly when Mg(2+) ions are in excess (50 mm). Without Mg(2+), IPP binds to UPPs only at high concentration. Mutation of Asp(26) to other charged amino acids results in significant decrease of the UPPs activity. The role of Asp(26) is probably to assist the migration of Mg(2+) from IPP to FPP and thus initiate the condensation reaction by ionization of the pyrophosphate group from FPP. Other conserved residues, including His(43), Ser(71), Asn(74), and Arg(77), may serve as general acid/base and pyrophosphate carrier. Our results here improve the understanding of the UPPs enzyme reaction significantly.  相似文献   

2.
The Escherichia coli undecaprayl-pyrophosphate synthase (UPPs) structure has been solved using the single wavelength anomalous diffraction method. The putative substrate-binding site is located near the end of the betaA-strand with Asp-26 playing a critical catalytic role. In both subunits, an elongated hydrophobic tunnel is found, surrounded by four beta-strands (betaA-betaB-betaD-betaC) and two helices (alpha2 and alpha3) and lined at the bottom with large residues Ile-62, Leu-137, Val-105, and His-103. The product distributions formed by the use of the I62A, V105A, and H103A mutants are similar to those observed for wild-type UPPs. Catalysis by the L137A UPPs, on the other hand, results in predominantly the formation of the C(70) polymer rather than the C(55) polymer. Ala-69 and Ala-143 are located near the top of the tunnel. In contrast to the A143V reaction, the C(30) intermediate is formed to a greater extent and is longer lived in the process catalyzed by the A69L mutant. These findings suggest that the small side chain of Ala-69 is required for rapid elongation to the C(55) product, whereas the large hydrophobic side chain of Leu-137 is required to limit the elongation to the C(55) product. The roles of residues located on a flexible loop were investigated. The S71A, N74A, or R77A mutants displayed 25-200-fold decrease in k(cat) values. W75A showed an 8-fold increase of the FPP K(m) value, and 22-33-fold increases in the IPP K(m) values were observed for E81A and S71A. The loop may function to bridge the interaction of IPP with FPP, needed to initiate the condensation reaction and serve as a hinge to control the substrate binding and product release.  相似文献   

3.
Pan JJ  Chiou ST  Liang PH 《Biochemistry》2000,39(35):10936-10942
Undecaprenyl pyrophosphate synthase (UPPs) catalyzes the condensation of eight molecules of isopentenyl pyrophosphate (IPP) with farnesyl pyrophosphate (FPP) to generate C(55) undecaprenyl pyrophosphate. We investigated the kinetics and mechanism of this reaction pathway using Escherichia coli UPPs. With a variety of different ratios of enzyme to substrate and FPP to IPP in the presence or absence of Triton, different product distributions were found. In the presence of excess FPP, the intermediates (C(25)-C(50)) accumulated. Under a condition with enzyme and FPP in excess of IPP, instead of C(20)-geranylgeranyl pyrophosphate, C(20), C(25), and C(30) were the major products. The UPPs steady-state k(cat) value (2.5 s(-1)) in the presence of 0.1% Triton was 190-fold larger than in the absence of Triton (0.013 s(-1)). The k(cat) value matched the rate constant of each IPP condensation obtained from the enzyme single-turnover experiments. This suggested that the IPP condensation rather than product release was the rate-limiting step in the presence of Triton. In the absence of Triton, the intermediates formed and disappeared in a similar manner under enzyme single turnover in contrast to the slow steady-state rate, which indicated a step after product generation was rate limiting. This was further supported by a burst product formation. Judging from the accumulation level of C(55), C(60), and C(65), their dissociation from the enzyme cannot be too slow and an even slower enzyme conformational change with a rate of 0.001 s(-1) might govern the UPPs reaction rate under the steady-state condition in the absence of Triton.  相似文献   

4.
Zheng R  Blanchard JS 《Biochemistry》2000,39(51):16244-16251
Ketopantoate reductase (EC 1.1.1.169) catalyzes the NADPH-dependent reduction of alpha-ketopantoate to D-(-)-pantoate in the biosynthesis of pantothenate. The pH dependence of V and V/K for the E. coli enzyme suggests the involvement of a general acid/base in the catalytic mechanism. To identify residues involved in catalysis and substrate binding, we mutated the following six strictly conserved residues to Ala: Lys72, Lys176, Glu210, Glu240, Asp248, and Glu256. Of these, the K176A and E256A mutant enzymes showed 233- and 42-fold decreases in V(max), and 336- and 63-fold increases in the K(m) value of ketopantoate, respectively, while the other mutants exhibited WT kinetic properties. The V(max) for the K176A and E256A mutant enzymes was markedly increased, up to 25% and 75% of the wild-type level, by exogenously added primary amines and formate, respectively. The rescue efficiencies for the K176A and E256A mutant enzymes were dependent on the molecular volume of rescue agents, as anticipated for a finite active site volume. The protonated form of the amine is responsible for recovery of activity, suggesting that Lys176 functions as a general acid in catalysis of ketopantoate reduction. The rescue efficiencies for the K176A mutant by primary amines were independent of the pK(a) value of the rescue agents (Bronsted coefficient, alpha = -0.004 +/-0.008). Insensitivity to acid strength suggests that the chemical reaction is not rate-limiting, consistent with (a) the catalytic efficiency of the wild-type enzyme (k(cat)/K(m) = 2x10(6) M(-1) s(-1) and (b) the small primary deuterium kinetic isotope effects, (D)V = 1.3 and (D)V/K = 1.5, observed for the wild-type enzyme. Larger primary deuterium isotope effects on V and V/K were observed for the K176A mutant ((D)V = 3.0, (D)V/K = 3.7) but decreased nearly to WT values as the concentration of ethylamine was increased. The nearly WT activity of the E256A mutant in the presence of formate argues for an important role for this residue in substrate binding. The double mutant (K176A/E256A) has no detectable ketopantoate reductase activity. These results indicate that Lys176 and Glu256 of the E. coli ketopantoate reductase are active site residues, and we propose specific roles for each in binding ketopantoate and catalysis.  相似文献   

5.
Yoon HY  Cho EH  Yang SJ  Lee HJ  Huh JW  Choi MM  Cho SW 《Biochimie》2004,86(4-5):261-267
In the present study, the cassette mutagenesis at several putative positions (K94, G96, K118, K130, or D172) was performed to examine the residues involved in the glutamate-binding of the human glutamate dehydrogenase isozymes (hGDH1 and hGDH2). None of the mutations tested affected the expression or stability of the proteins. There was dramatic reduction in the catalytic efficiency in mutant proteins at K94, G96, K118, or K130 site, but not at D172 site. The K(M) values for glutamate were 4-10-fold greater for the mutants at K94, G96, or K118 site than for the wild-type hGDH1 and hGDH2, whereas no differences in the K(M) values for NAD(+) were detected between the mutant and wild-type enzymes. For K130Y mutant, the K(M) value for glutamate increased 1.6-fold, whereas the catalytic efficiency (k(cat)/K(M)) showed only 2-3% of the wild-type. Therefore, the decreased catalytic efficiency of the K130 mutant mainly results from the reduced k(cat) value, suggesting a possibility that the K130Y residue may be involved in the catalysis rather than in the glutamate-binding. The D172Y mutant did not show any changes in k(cat) value and K(M) values for glutamate and NAD(+), indicating that D172Y is not directly involved in catalysis and substrates binding of the hGDH isozymes. For sensitivity to ADP activation, only the D172Y mutant showed a reduced sensitivity to ADP activation. The reduction of ADP activation in D172Y mutant was more profoundly observed in hGDH2 than in hGDH1. There were no differences in their sensitivities to GTP inhibition between the wild-type and mutant GDHs at all positions tested. Our results suggest that K94, G96, and K118 residues play an important role, although at different degrees, in the binding of glutamate to hGDH isozymes.  相似文献   

6.
Undecaprenyl pyrophosphate synthase (UPPs) catalyzes chain elongation of farnesyl pyrophosphate (FPP) to undecaprenyl pyrophosphate (UPP) via condensation with eight isopentenyl pyrophosphates (IPP). UPPs from Escherichia coli is a dimer, and each subunit consists of 253 amino acid residues. The chain length of the product is modulated by a hydrophobic active site tunnel. In this paper, the crystal structure of E. coli UPPs was refined to 1.73 A resolution, which showed bound sulfate and magnesium ions as well as Triton X-100 molecules. The amino acid residues 72-82, which encompass an essential catalytic loop not seen in the previous apoenzyme structure (Ko, T.-P., Chen, Y. K., Robinson, H., Tsai, P. C., Gao, Y.-G., Chen, A. P.-C., Wang, A. H.-J., and Liang, P.-H. (2001) J. Biol. Chem. 276, 47474-47482), also became visible in one subunit. The sulfate ions suggest locations of the pyrophosphate groups of FPP and IPP in the active site. The Mg2+ is chelated by His-199 and Glu-213 from different subunits and possibly plays a structural rather than catalytic role. However, the metal ion is near the IPP-binding site, and double mutation of His-199 and Glu-213 to alanines showed a remarkable increase of Km value for IPP. Inside the tunnel, one Triton surrounds the top portion of the tunnel, and the other occupies the bottom part. These two Triton molecules may mimic the hydrocarbon moiety of the UPP product in the active site. Kinetic analysis indicated that a high concentration (>1%) of Triton inhibits the enzyme activity.  相似文献   

7.
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   

8.
We have performed random mutagenesis coupled with selection to isolate mutant enzymes with high catalytic activities at low temperature from thermophilic 3-isopropylmalate dehydrogenase (IPMDH) originally isolated from Thermus thermophilus. Five cold-adapted mutant IPMDHs with single-amino-acid substitutions were obtained and analyzed. Kinetic analysis revealed that there are two types of cold-adapted mutant IPMDH: k(cat)-improved (improved in k(cat)) and K(m)-improved (improved in k(cat)/K(m)) types. To determine the mechanisms of cold adaptation of these mutants, thermodynamic parameters were estimated and compared with those of the Escherichia coli wild-type IPMDH. The Delta G(m) values for Michaelis intermediate formation of the k(cat)-improved-type enzymes were larger than that of the T. thermophilus wild-type IPMDH and similar to that of the E. coli wild-type IPMDH. The Delta G(m) values of K(m)-improved-type enzymes were smaller than that of the T. thermophilus wild-type IPMDH. Fitting of NAD(+) binding was improved in the K(m)-improved-type enzymes. The two types of cold-adapted mutants employed one of the two strategies of E. coli wild-type IPMDH: relative destabilization of the Michaelis complex in k(cat)-improved-type, and destabilization of the rate-limiting step in K(m)-improved type mutants. Some cold-adapted mutant IPMDHs retained thermostability similar to that of the T. thermophilus wild-type IPMDH.  相似文献   

9.
The functional and structural significance of amino acid residues Met(39), Glu(56), Asp(58), Glu(60), and Gly(63) of Fibrobacter succinogenes 1,3-1,4-beta-d-glucanase was explored by the approach of site-directed mutagenesis, initial rate kinetics, fluorescence spectroscopy, and CD spectrometry. Glu(56), Asp(58), Glu(60), and Gly(63) residues are conserved among known primary sequences of the bacterial and fungal enzymes. Kinetic analyses revealed that 240-, 540-, 570-, and 880-fold decreases in k(cat) were observed for the E56D, E60D, D58N, and D58E mutant enzymes, respectively, with a similar substrate affinity relative to the wild type enzyme. In contrast, no detectable enzymatic activity was observed for the E56A, E56Q, D58A, E60A, and E60Q mutants. These results indicated that the carboxyl side chain at positions 56 and 60 is mandatory for enzyme catalysis. M39F, unlike the other mutants, exhibited a 5-fold increase in K(m) value. Lower thermostability was found with the G63A mutant when compared with wild type or other mutant forms of F. succinogenes 1,3-1,4-beta-d-glucanase. Denatured wild type and mutant enzymes were, however, recoverable as active enzymes when 8 m urea was employed as the denaturant. Structural modeling and kinetic studies suggest that Glu(56), Asp(58), and Glu(60) residues apparently play important role(s) in the catalysis of F. succinogenes 1,3-1,4-beta-d-glucanase.  相似文献   

10.
The catalytic reaction mechanism and binding of substrates was investigated for the multisubstrate Drosophila melanogaster deoxyribonucleoside kinase. Mutation of E52 to D, Q and H plus mutations of R105 to K and H were performed to investigate the proposed catalytic reaction mechanism, in which E52 acts as an initiating base and R105 is thought to stabilize the transition state of the reaction. Mutant enzymes (E52D, E52H and R105H) showed a markedly decreased k(cat), while the catalytic activity of E52Q and R105K was abolished. The E52D mutant was crystallized with its feedback inhibitor dTTP. The backbone conformation remained unchanged, and coordination between D52 and the dTTP-Mg complex was observed. The observed decrease in k(cat) for E52D was most likely due to an increased distance between the catalytic carboxyl group and 5'-OH of deoxythymidine (dThd) or deoxycytidine (dCyd). Mutation of Q81 to N and Y70 to W was carried out to investigate substrate binding. The mutations primarily affected the K(m) values, whereas the k(cat) values were of the same magnitude as for the wild-type. The Y70W mutation made the enzyme lose activity towards purines and negative cooperativity towards dThd and dCyd was observed. The Q81N mutation showed a 200- and 100-fold increase in K(m), whereas k(cat) was decreased five- and twofold for dThd and dCyd, respectively, supporting a role in substrate binding. These observations give insight into the mechanisms of substrate binding and catalysis, which is important for developing novel suicide genes and drugs for use in gene therapy.  相似文献   

11.
Hirano N  Haruki M  Morikawa M  Kanaya S 《Biochemistry》2000,39(43):13285-13294
A genetic method for isolating a mutant enzyme of ribonuclease HI (RNase HI) from Thermus thermophilus HB8 with enhanced activity at moderate temperatures was developed. T. thermophilus RNase HI has an ability to complement the RNase H-dependent temperature-sensitive (ts) growth phenotype of Escherichia coli MIC3001. However, this complementation ability was greatly reduced by replacing Asp(134), which is one of the active site residues, with His, probably due to a reduction in the catalytic activity. Random mutagenesis of the gene encoding the resultant D134H enzyme, followed by screening for second-site revertants, allowed us to isolate three single mutations (Ala(12) --> Ser, Lys(75) --> Met, and Ala(77) --> Pro) that restore the normal complementation ability to the D134H enzyme. These mutations were individually or simultaneously introduced into the wild-type enzyme, and the kinetic parameters of the resultant mutant enzymes for the hydrolysis of a DNA-RNA-DNA/DNA substrate were determined at 30 degrees C. Each mutation increased the k(cat)/K(m) value of the wild-type enzyme by 2.1-4.8-fold. The effects of the mutations on the enzymatic activity were roughly cumulative, and the combination of these three mutations increased the k(cat)/K(m) value of the wild-type enzyme by 40-fold (5.5-fold in k(cat)). Measurement of thermal stability of the mutant enzymes with circular dichroism spectroscopy in the presence of 1 M guanidine hydrochloride and 1 mM dithiothreitol showed that the T(m) value of the triple mutant enzyme, in which all three mutations were combined, was comparable to that of the wild-type enzyme (75.0 vs 77.4 degrees C). These results demonstrate that the activity of a thermophilic enzyme can be improved without a cost of protein stability.  相似文献   

12.
GDP-mannose hydrolase (GDPMH) catalyzes the hydrolysis of GDP-alpha-d-sugars by nucleophilic substitution with inversion at the anomeric C1 atom of the sugar, with general base catalysis by H124. Three lines of evidence indicate a mechanism with dissociative character. First, in the 1.3 A X-ray structure of the GDPMH-Mg(2+)-GDP.Tris(+) complex [Gabelli, S. B., et al. (2004) Structure 12, 927-935], the GDP leaving group interacts with five catalytic components: R37, Y103, R52, R65, and the essential Mg(2+). As determined by the effects of site-specific mutants on k(cat), these components contribute factors of 24-, 100-, 309-, 24-, and >/=10(5)-fold, respectively, to catalysis. Both R37 and Y103 bind the beta-phosphate of GDP and are only 5.0 A apart. Accordingly, the R37Q/Y103F double mutant exhibits partially additive effects of the two single mutants on k(cat), indicating cooperativity of R37 and Y103 in promoting catalysis, and antagonistic effects on K(m). Second, the conserved residue, D22, is positioned to accept a hydrogen bond from the C2-OH group of the sugar undergoing substitution at C1, as was shown by modeling an alpha-d-mannosyl group into the sugar binding site. The D22A and D22N mutations decreased k(cat) by factors of 10(2.1) and 10(2.6), respectively, for the hydrolysis of GDP-alpha-d-mannose, and showed smaller effects on K(m), suggesting that the D22 anion stabilizes a cationic oxocarbenium transition state. Third, the fluorinated substrate, GDP-2F-alpha-d-mannose, for which a cationic oxocarbenium transition state would be destabilized by electron withdrawal, exhibited a 16-fold decrease in k(cat) and a smaller, 2.5-fold increase in K(m). The D22A and D22N mutations further decreased the k(cat) with GDP-2F-alpha-d-mannose to values similar to those found with GDP-alpha-d-mannose, and decreased the K(m) of the fluorinated substrate. The choice of histidine as the general base over glutamate, the preferred base in other Nudix enzymes, is not due to the greater basicity of histidine, since the pK(a) of E124 in the active complex (7.7) exceeded that of H124 (6.7), and the H124E mutation showed a 10(2.2)-fold decrease in k(cat) and a 4.0-fold increase in K(m) at pH 9.3. Similarly, the catalytic triad detected in the X-ray structure (H124- - -Y127- - -P120) is unnecessary for orienting H124, since the Y127F mutation had only 2-fold effects on k(cat) and K(m) with either H124 or E124 as the general base. Hence, a neutral histidine rather than an anionic glutamate may be necessary to preserve electroneutrality in the active complex.  相似文献   

13.
The role of Ser 167 of Escherichia coli thymidylate synthase (TS) in catalysis has been characterized by kinetic and crystallographic studies. Position 167 variants including S167A, S167N, S167D, S167C, S167G, S167L, S167T, and S167V were generated by site-directed mutagenesis. Only S167A, S167G, S167T, and S167C complemented the growth of thymidine auxotrophs of E. coli in medium lacking thymidine. Steady-state kinetic analysis revealed that mutant enzymes exhibited k(cat) values 1.1-95-fold lower than that of the wild-type enzyme. Relative to wild-type TS, K(m) values of the mutant enzymes for 2'-deoxyuridylate (dUMP) were 5-90 times higher, while K(m) values for 5,10-methylenetetrahydrofolate (CH(2)H(4)folate) were 1.5-16-fold higher. The rate of dehalogenation of 5-bromo-2'-deoxyuridine 5'-monophosphate (BrdUMP), a reaction catalyzed by TS that does not require CH(2)H(4)folate as cosubstrate, by mutant TSs was analyzed and showed that only S167A and S167G catalyzed the dehalogenation reaction and values of k(cat)/K(m) for the mutant enzymes were decreased by 10- and 3000-fold, respectively. Analysis of pre-steady-state kinetics of ternary complex formation revealed that the productive binding of CH(2)H(4)folate is weaker to mutant TSs than to the wild-type enzyme. Chemical transformation constants (k(chem)) for the mutant enzymes were lower by 1.1-6.0-fold relative to the wild-type enzyme. S167A, S167T, and S167C crystallized in the I2(1)3 space group and scattered X-rays to either 1.7 A (S167A and S167T) or 2.6 A (S167C). The high-resolution data sets were refined to a R(crys) of 19.9%. In the crystals some cysteine residues were derivatized with 2-mercaptoethanol to form S,S-(2-hydroxyethyl)thiocysteine. The pattern of derivatization indicates that in the absence of bound substrate the catalytic cysteine is not more reactive than other cysteines. It is proposed that the catalytic cysteine is activated by substrate binding by a proton-transfer mechanism in which the phosphate group of the nucleotide neutralizes the charge of Arg 126', facilitating the transfer of a proton from the catalytic cysteine to a His 207-Asp 205 diad via a system of ordered water molecules.  相似文献   

14.
Phosphite dehydrogenase (PTDH) from Pseudomonas stutzeri catalyzes the nicotinamide adenine dinucleotide-dependent oxidation of phosphite to phosphate. The enzyme belongs to the family of D-hydroxy acid dehydrogenases (DHDHs). A search of the protein databases uncovered many additional putative phosphite dehydrogenases. The genes encoding four diverse candidates were cloned and expressed, and the enzymes were purified and characterized. All oxidized phosphite to phosphate and had similar kinetic parameters despite a low level of pairwise sequence identity (39-72%). A recent crystal structure identified Arg301 as a residue in the active site that has not been investigated previously. Arg301 is fully conserved in the enzymes shown here to be PTDHs, but the residue is not conserved in other DHDHs. Kinetic analysis of site-directed mutants of this residue shows that it is important for efficient catalysis, with an ~100-fold decrease in k(cat) and an almost 700-fold increase in K(m,phosphite) for the R301A mutant. Interestingly, the R301K mutant displayed a slightly higher k(cat) than the parent PTDH, and a more modest increase in K(m) for phosphite (nearly 40-fold). Given these results, Arg301 may be involved in the binding and orientation of the phosphite substrate and/or play a catalytic role via electrostatic interactions. Three other residues in the active site region that are conserved in the PTDH orthologs but not DHDHs were identified (Trp134, Tyr139, and Ser295). The importance of these residues was also investigated by site-directed mutagenesis. All of the mutants had k(cat) values similar to that of the wild-type enzyme, indicating these residues are not important for catalysis.  相似文献   

15.
The roles of particular amino acids in substrate and coenzyme binding and catalysis of glucose-6-phosphate dehydrogenase of Leuconostoc mesenteroides have been investigated by site-directed mutagenesis, kinetic analysis, and determination of binding constants. The enzyme from this species has functional dual NADP(+)/NAD(+) specificity. Previous investigations in our laboratories determined the three-dimensional structure. Kinetic studies showed an ordered mechanism for the NADP-linked reaction while the NAD-linked reaction is random. His-240 was identified as the catalytic base, and Arg-46 was identified as important for NADP(+) but not NAD(+) binding. Mutations have been selected on the basis of the three-dimensional structure. Kinetic studies of 14 mutant enzymes are reported and kinetic mechanisms are reported for 5 mutant enzymes. Fourteen substrate or coenzyme dissociation constants have been measured for 11 mutant enzymes. Roles of particular residues are inferred from k(cat), K(m), k(cat)/K(m), K(d), and changes in kinetic mechanism. Results for enzymes K182R, K182Q, K343R, and K343Q establish Lys-182 and Lys-343 as important in binding substrate both to free enzyme and during catalysis. Studies of mutant enzymes Y415F and Y179F showed no significant contribution for Tyr-415 to substrate binding and only a small contribution for Tyr-179. Changes in kinetics for T14A, Q47E, and R46A enzymes implicate these residues, to differing extents, in coenzyme binding and discrimination between NADP(+) and NAD(+). By the same measure, Lys-343 is also involved in defining coenzyme specificity. Decrease in k(cat) and k(cat)/K(m) for the D374Q mutant enzyme defines the way Asp-374, unique to L. mesenteroides G6PD, modulates stabilization of the enzyme during catalysis by its interaction with Lys-182. The greatly reduced k(cat) values of enzymes P149V and P149G indicate the importance of the cis conformation of Pro-149 in accessing the correct transition state.  相似文献   

16.
Glutathione is essential for maintaining the intracellular redox environment and is synthesized from gamma-glutamylcysteine, glycine, and ATP by glutathione synthetase (GS). To examine the reaction mechanism of a eukaryotic GS, 24 Arabidopsis thaliana GS (AtGS) mutants were kinetically characterized. Within the gamma-glutamylcysteine/glutathione-binding site, the S153A and S155A mutants displayed less than 4-fold changes in kinetic parameters with mutations of Glu-220 (E220A/E220Q), Gln-226 (Q226A/Q226N), and Arg-274 (R274A/R274K) at the distal end of the binding site resulting in 24-180-fold increases in the K(m) values for gamma-glutamylcysteine. Substitution of multiple residues interacting with ATP (K313M, K367M, and E429A/E429Q) or coordinating magnesium ions to ATP (E148A/E148Q, N150A/N150D, and E371A) yielded inactive protein because of compromised nucleotide binding, as determined by fluorescence titration. Other mutations in the ATP-binding site (E371Q, N376A, and K456M) resulted in greater than 30-fold decreases in affinity for ATP and up to 80-fold reductions in turnover rate. Mutation of Arg-132 and Arg-454, which are positioned at the interface of the two substrate-binding sites, affected the enzymatic activity differently. The R132A mutant was inactive, and the R132K mutant decreased k(cat) by 200-fold; however, both mutants bound ATP with K(d) values similar to wild-type enzyme. Minimal changes in kinetic parameters were observed with the R454K mutant, but the R454A mutant displayed a 160-fold decrease in k(cat). In addition, the R132K, R454A, and R454K mutations elevated the K(m) value for glycine up to 11-fold. Comparison of the pH profiles and the solvent deuterium isotope effects of A. thaliana GS and the Arg-132 and Arg-454 mutants also suggest distinct mechanistic roles for these residues. Based on these results, a catalytic mechanism for the eukaryotic GS is proposed.  相似文献   

17.
Alanine-scanning site-directed mutagenesis was carried out on all conserved lysine residues of pigeon cytosolic NADP(+)-dependent malic enzyme. Only two mutant enzymes, K162A and K340A, showed significant effect on their kinetic parameters. Both mutant enzymes have K(m) values for Mn(2+) and l-malate similar to those of wild-type. The K(m) value for NADP(+) of K162A is identical to that of wild-type. However, K162A demonstrated a 235-fold decrease in the k(cat) value (0.17 +/- 0.01 vs 40.0 +/- 1.3 s(-1)). These data suggested that the side chain of K162 is important for the enzyme catalytic reaction. We propose that the epsilon-amino group of K162 may serve as a general acid to protonate the 3-carbon of enolpyruvate after decarboxylation. The K340A mutant demonstrated no effect on the k(cat) value. However, its K(m) value for NADP(+) was increased by a factor of 65 (225.7 +/- 5.07 vs 3.49 +/- 0.05 microM). We propose that the NADP(+) specificity is determined by the electrostatic interaction between the epsilon-amino group of K340 and 2'-phosphate of NADP(+).  相似文献   

18.
Fatty acid amide hydrolase (FAAH) is a membrane-bound enzyme responsible for the catabolism of neuromodulatory fatty acid amides, including anandamide and oleamide. FAAH's primary structure identifies this enzyme as a member of a diverse group of alkyl amidases, known collectively as the "amidase signature family". At present, this enzyme family's catalytic mechanism remains poorly understood. In this study, we investigated the catalytic features of FAAH through mutagenesis, affinity labeling, and steady-state kinetic methods. In particular, we focused on the respective roles of three serine residues that are conserved in all amidase signature enzymes (S217, S218, and S241 in FAAH). Mutation of each of these serines to alanine resulted in a FAAH enzyme bearing significant catalytic defects, with the S217A and S218A mutants showing 2300- and 95-fold reductions in k(cat), respectively, and the S241A mutant exhibiting no detectable catalytic activity. The double S217A:S218A FAAH mutant displayed a 230 000-fold decrease in k(cat), supporting independent catalytic functions for these serine residues. Affinity labeling of FAAH with a specific nucleophile reactive inhibitor, ethoxy oleoyl fluorophosphonate, identified S241 as the enzyme's catalytic nucleophile. The pH dependence of FAAH's k(cat) and k(cat)/K(m) implicated a base involved in catalysis with a pK(a) of 7.9. Interestingly, mutation of each of FAAH's conserved histidines (H184, H358, and H449) generated active enzymes, indicating that FAAH does not contain a Ser-His-Asp catalytic triad commonly found in other mammalian serine hydrolytic enzymes. The unusual properties of FAAH identified here suggest that this enzyme, and possibly the amidase signature family as a whole, may hydrolyze amides by a novel catalytic mechanism.  相似文献   

19.
The roles of three conserved active site carboxylic acids (D197, E233, and D300) in the catalytic mechanism of human pancreatic alpha-amylase (HPA) were studied by utilizing site-directed mutagenesis in combination with structural and kinetic analyses of the resultant enzymes. All three residues were mutated to both alanine and the respective amide, and a double alanine mutant (E233A/D300A) was also generated. Structural analyses demonstrated that there were no significant differences in global fold for the mutant enzymes. Kinetic analyses were performed on the mutants, utilizing a range of substrates. All results suggested that D197 was the nucleophile, as virtually all activity (>10(5)-fold decrease in k(cat) values) was lost for the enzymes mutated at this position when assayed with several substrates. The significantly greater second-order rate constant of E233 mutants on "activated" substrates (k(cat)/K(m) value for alpha-maltotriosyl fluoride = 15 s(-)(1) mM(-)(1)) compared with "unactivated" substrates (k(cat)/K(m) value for maltopentaose = 0.0030 s(-)(1) mM(-)(1)) strongly suggested that E233 is the general acid catalyst, as did the pH-activity profiles. Transglycosylation was favored over hydrolysis for the reactions of several of the enzymes mutated at D300. At the least, this suggests an overall impairment of the catalytic mechanism where the reaction then proceeds using the better acceptor (oligosaccharide instead of water). This may also suggest that D300 plays a crucial role in enzymic interactions with the nucleophilic water during the hydrolysis of the glycosidic bond.  相似文献   

20.
1-Deoxy-d-xylulose 5-phosphate (DXP) reductoisomerase, which simultaneously catalyzes the intramolecular rearrangement and reduction of DXP to form 2-C-methyl-d-erythritol 4-phosphate, constitutes a key enzyme of an alternative mevalonate-independent pathway for isopentenyl diphosphate biosynthesis. The dxr gene encoding this enzyme from Escherichia coli was overexpressed as a histidine-tagged protein and characterized in detail. DNA sequencing analysis of the dxr genes from 10 E. coli dxr-deficient mutants revealed base substitution mutations at four points: two nonsense mutations and two amino acid substitutions (Gly(14) to Asp(14) and Glu(231) to Lys(231)). Diethyl pyrocarbonate treatment inactivated DXP reductoisomerase, and subsequent hydroxylamine treatment restored the activity of the diethyl pyrocarbonate-treated enzyme. To characterize these defects, we overexpressed the mutant enzymes G14D, E231K, H153Q, H209Q, and H257Q. All of these mutant enzymes except for G14D were obtained as soluble proteins. Although the purified enzyme E231K had wild-type K(m) values for DXP and NADPH, the mutant enzyme had less than a 0.24% wild-type k(cat) value. K(m) values of H153Q, H209Q, and H257Q for DXP increased to 3.5-, 7.6-, and 19-fold the wild-type value, respectively. These results indicate that Glu(231) of E. coli DXP reductoisomerase plays an important role(s) in the conversion of DXP to 2-C-methyl-d-erythritol 4-phosphate, and that His(153), His(209), and His(257), in part, associate with DXP binding in the enzyme molecule.  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