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 0 毫秒
1.
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   

2.
Zheng R  Blanchard JS 《Biochemistry》2003,42(38):11289-11296
Ketopantoate reductase (EC 1.1.1.169), an enzyme in the pantothenate biosynthetic pathway, catalyzes the NADPH-dependent reduction of alpha-ketopantoate to form D-(-)-pantoate. The enzyme exhibits high specificity for ketopantoate, with V and V/K for ketopantoate being 5- and 365-fold higher than those values for alpha-ketoisovalerate and 20- and 648-fold higher than those values for alpha-keto-beta-methyl-n-valerate, respectively. For pyridine nucleotides, V/K for beta-NADPH is 3-500-fold higher than that for other nucleotide substrates. The magnitude of the primary deuterium kinetic isotope effects on V and V/K varied substantially when different ketoacid and pyridine nucleotide substrates were used. The small primary deuterium kinetic isotope effects observed using NADPH and NHDPH suggest that the chemical step is not rate-limiting, while larger primary deuterium isotope effects were observed for poor ketoacid and pyridine nucleotide substrates, indicating that the chemical reaction has become partially or completely rate-limiting. The pH dependence of (D)V using ketopantoate was observed to vary from a value of 1.1 at low pH to a value of 2.5 at high pH, while the magnitude of (D)V/K(NADPH) and (D)V/K(KP) were pH-independent. The value of (D)V is large and pH-independent when alpha-keto-beta-methyl-n-valerate was used as the ketoacid substrate. Solvent kinetic isotope effects of 2.2 and 1.2 on V and V/K, respectively, were observed with alpha-keto-beta-methyl-n-valerate. Rapid reaction analysis of NADPH oxidation using ketopantoate showed no "burst" phase, suggesting that product-release steps are not rate-limiting and the cause of the small observed kinetic isotope effects with this substrate pair. Large primary deuterium isotope effects on V and V/K using 3-APADPH in steady-state experiments, equivalent to the isotope effect observed in single turnover studies, suggests that chemistry is rate-limiting for this poorer reductant. These results are discussed in terms of a kinetic and chemical mechanism for the enzyme.  相似文献   

3.
Pollock VV  Barber MJ 《Biochemistry》2001,40(5):1430-1440
Rhodobacter sphaeroides f. sp. denitrificans biotin sulfoxide reductase catalyzes the reduction of d-biotin d-sulfoxide (BSO) to biotin. Initial rate studies of the homogeneous recombinant enzyme, expressed in Escherichia coli, have demonstrated that the purified protein utilizes NADPH as a facile electron donor in the absence of any additional auxiliary proteins. We have previously shown [Pollock, V. V., and Barber, M. J. (1997) J. Biol. Chem. 272, 3355-3362] that, at pH 8 and in the presence of saturating concentrations of BSO, the enzyme exhibits, a marked preference for NADPH (k(cat,app) = 500 s(-1), K(m,app) = 269 microM, and k(cat,app)/K(m,app) = 1.86 x 10(6) M(-1) s(-1)) compared to NADH (k(cat,app) = 47 s(-1), K(m,app) = 394 microM, and k(cat,app)/K(m,app) = 1.19 x 10(5) M(-1) s(-1)). Production of biotin using NADPH as the electron donor was confirmed by both the disk biological assay and by reversed-phase HPLC analysis of the reaction products. The purified enzyme also utilized ferricyanide as an artificial electron acceptor, which effectively suppressed biotin sulfoxide reduction and biotin formation. Analysis of the enzyme isolated from tungsten-grown cells yielded decreased reduced methyl viologen:BSO reductase, NADPH:BSO reductase, and NADPH:FR activities, confirming that Mo is required for all activities. Kinetic analyses of substrate inhibition profiles revealed that the enzyme followed a Ping Pong Bi-Bi mechanism with both NADPH and BSO exhibiting double competitive substrate inhibition. Replots of the 1/v-axes intercepts of the parallel asymptotes obtained at several low concentrations of fixed substrate yielded a K(m) for BSO of 714 and 65 microM for NADPH. In contrast, utilizing NADH as an electron donor, the replots yielded a K(m) for BSO of 132 microM and 1.25 mM for NADH. Slope replots of data obtained at high concentrations of BSO yielded a K(i) for BSO of 6.10 mM and 900 microM for NADPH. Kinetic isotope studies utilizing stereospecifically deuterated NADPD indicated that BSO reductase uses specifically the 4R-hydrogen of the nicotinamide ring. Cyanide inhibited NADPH:BSO and NADPH:FR activities in a reversible manner while diethylpyrocarbonate treatment resulted in complete irreversible inactivation of the enzyme concomitant with molybdenum cofactor release, indicating that histidine residues are involved in cofactor-binding.  相似文献   

4.
A kinetic mechanism is presented for Escherichia coli dihydrofolate reductase which describes the full time course of the enzymatic reaction over a wide range of substrate and enzyme concentrations at pH 7.2 and 20 degrees C. Specific rate constants were estimated by computer simulation of the full time course of single turnover, burst, and steady-state experiments using both nondeuterated and deuterated NADPH. The mechanism involves the random addition of substrates, but the substrates and enzyme are not at equilibrium prior to the chemical transformation step. The rate-limiting step follows the chemical transformation, and the maximum velocity of the reaction is limited by the release of the product tetrahydrofolate. The full time course of the reaction is markedly affected by the formation of the enzyme-NADPH-tetrahydrofolate abortive complex, but not by the enzyme-NADP-dihydrofolate abortive complex.  相似文献   

5.
The NADPH-dependent reduction of ketopantoate to pantoate, catalyzed by ketopantoate reductase (KPR; EC 1.1.1.169), is essential for the biosynthesis of pantothenate (vitamin B(5)). Here we present the crystal structure of Escherichia coli KPR with NADP(+) bound, solved to 2.1 A resolution. The cofactor is bound in the active site cleft between the N-terminal Rossmann-fold domain and the C-terminal alpha-helical domain. The thermodynamics of cofactor and substrate binding were characterized by isothermal titration calorimetry. The dissociation constant for NADP(+) was found to be 6.5 muM, 20-fold larger than that for NADPH (0.34 muM). The difference is primarily due to the entropic term, suggesting favorable hydrophobic interactions of the more lipophilic nicotinamide ring in NADPH. Comparison of this binary complex structure with the previously studied apoenzyme reveals no evidence for large domain movements on cofactor binding. This observation is further supported both by molecular dynamics and by calorimetric analysis. A model of the ternary complex, based on the structure presented here, provides novel insights into the molecular mechanism of enzyme catalysis. We propose a conformational switch of the essential Lys176 from the "resting" state observed in our structure to an "active" state, to bind ketopantoate. Additionally, we identify the importance of Asn98 for substrate binding and enzyme catalysis.  相似文献   

6.
We report the crystal structure of E. coli ketopantoate hydroxymethyltransferase (KPHMT) at 1.9 A resolution, in complex with its product, ketopantoate. KPHMT catalyzes the first step in the biosynthesis of pantothenate (vitamin B(5)), the precursor of coenzyme A and the acyl carrier protein cofactor. The structure of the decameric enzyme was solved by multiwavelength anomalous dispersion to locate 160 selenomethionine sites and phase 560 kDa of protein, making it the largest structure solved by this approach. KPHMT adopts the (betaalpha)(8) barrel fold and is a member of the phosphoenolpyruvate/pyruvate superfamily. The active site contains a ketopantoate bidentately coordinated to Mg(2+). Similar binding is likely for the substrate, alpha-ketoisovalerate, orienting the C3 for deprotonation.  相似文献   

7.
A mutation in the pheS gene, encoding phenylalanyl-tRNA synthetase, in E. coli NP37 confers temperature-sensitivity on the organism. A five-fold increase in tRNA(phe) levels complements the mutation. Analysis of the kinetic properties of the mutant enzyme indicates that the KM is 20-fold higher than the wild-type and the dissociation constant of the tRNA(phe)-synthetase complex for the mutant is at least 10-fold higher. These results indicate that the mutation in E. coli NP37 directly affects the tRNA(phe) binding site on the cognate synthetase.  相似文献   

8.
Ketopantoate reductase (KPR, EC 1.1.1.169) catalyzes the NADPH-dependent reduction of ketopantoate to pantoate on the pantothenate (vitamin B(5)) biosynthetic pathway. The Escherichia coli panE gene encoding KPR was cloned and expressed at high levels as the native and selenomethionine-substituted (SeMet) proteins. Both native and SeMet recombinant proteins were purified by three chromatographic steps, to yield pure proteins. The wild-type enzyme was found to have a K(M)(NADPH) of 20 microM, a K(M)(ketopantoate) of 60 microM, and a k(cat) of 40 s(-1). Regular prismatic KPR crystals were prepared using the hanging drop technique. They belonged to the tetragonal space group P4(2)2(1)2, with cell parameters: a = b = 103.7 A and c = 55.7 A, accommodating one enzyme molecule per asymmetric unit. The structure of KPR was determined by the multiwavelength anomalous dispersion method using the SeMet protein, for which data were collected to 2.3 A resolution. The native data were collected to 1.7 A resolution and used to refine the final structure. The secondary structure comprises 12 alpha-helices, three 3(10)-helices, and 11 beta-strands. The enzyme is monomeric and has two domains separated by a cleft. The N-terminal domain has an alphabeta-fold of the Rossmann type. The C-terminal domain (residues 170-291) is composed of eight alpha-helices. KPR is shown to be a member of the 6-phosphogluconate dehydrogenase C-terminal domain-like superfamily. A model for the ternary enzyme-NADPH-ketopantoate ternary complex provides a rationale for kinetic data reported for specific site-directed mutants.  相似文献   

9.
Using gene replacement and transposon Tn5 mutagenesis, an Escherichia coli ilvC panE double mutant completely lacking ketopantoate reductase activity was isolated. This E. coli double mutant was employed to isolate the E. coli panE gene by genetic complementation. The E. coli panE gene is characterized by a 912 bp coding region, which specifies a protein of 303 amino acids with a deduced molecular mass of 33.8 kD. A panE expression plasmid carrying the panE gene under the control of the tac promotor was constructed. Introduction of the panE expression plasmid into E. coli resulted in a threefold increase in ketopantoate reductase activity. It was also shown that the enhanced panE expression in E. coli K12 led to 3.5-fold increase in pantothenate excretion. Pantothenate excretion could even be more enhanced when the growth medium was supplemented with ketopantoate.  相似文献   

10.
Further to a linear free-energy correlation of cross-reaction rate constants k12 for the reaction of eight organic radicals (OR), e.g. MV*+, from methyl viologen, with cytochrome c(III), we consider here similar studies for the reduction of the R2 protein of Escherichia coli ribonucleotide reductase, which has FeIII2 and Tyr* redox components. The same two techniques of pulse radiolysis and stopped-flow were used. Cross-reaction rate constants (22 degrees C) at pH 7.0, I=0.100 M (NaCl), were determined for the reduction of active-R2 with the eight ORs, reduction potentials E0(1) from -0.446 to +0.194 V. Samples of active-R2 have an FeIII2 met-R2 component, which in the present studies was close to 40%. Concurrent reactions have to be taken into account for the five most reactive ORs, corresponding to reduction of the FeIII2 of met-R2 and then of active-R2. Separate experiments on met-R2 reproduced the first of these rate constants, which on average is approximately 66% larger than the second rate constant. A single Marcus free-energy plot of log k12-0.5 log10f versus -E0(1)/0.059 describes all the data and the slope of 0.54 is in satisfactory agreement with the theoretical value of 0.50. Such behaviour is unexpected since the Tyr* is a much stronger oxidant (E0 approximately 1.0 V versus NHE) as compared to FeIII2 (E0 close to zero). X-ray structures of the met- and red-R2 states have indicated that electroneutrality of the approximately 10 A buried active site is maintained. Proton transfer is therefore proposed as a rapid sequel to electron transfer. Other reactions considered are the much slower conventional time-range reductions of active-R2 with hydrazine and dithionite. For these reactions one and/or two-equivalent changes are possible. With both reductants, met-R2 reacts about four-fold faster than active-R2, and as with the ORs the less strongly oxidising FeIII2 component is reduced before the Tyr*.  相似文献   

11.
Riboflavin is biosynthesized by most microorganisms and plants, while mammals depend entirely on the absorption of this vitamin from the diet to meet their metabolic needs. Therefore, riboflavin biosynthesis appears to be an attractive target for drug design, since appropriate inhibitors of the pathway would selectively target the microorganism. We have cloned and solubly expressed the bifunctional ribD gene from Escherichia coli, whose three-dimensional structure was recently determined. We have demonstrated that the rate of deamination (370 min (-1)) exceeds the rate of reduction (19 min (-1)), suggesting no channeling between the two active sites. The reductive ring opening reaction occurs via a hydride transfer from the C 4- pro-R hydrogen of NADPH to C'-1 of ribose and is the rate-limiting step in the overall reaction, exhibiting a primary kinetic isotope effect ( (D) V) of 2.2. We also show that the INH-NADP adduct, one of the active forms of the anti-TB drug isoniazid, inhibits the E. coli RibD. On the basis of the observed patterns of inhibition versus the two substrates, we propose that the RibD-catalyzed reduction step follows a kinetic scheme similar to that of its structural homologue, DHFR.  相似文献   

12.
Ge J  Yu G  Ator MA  Stubbe J 《Biochemistry》2003,42(34):10071-10083
E. coli ribonucleotide reductase (RNR) catalyzes the conversion of nucleoside diphosphates (NDPs) to dNDPs and is composed of two homodimeric subunits: R1 and R2. R1 binds NDPs and contains binding sites for allosteric effectors that control substrate specificity and turnover rate. R2 contains a diiron-tyrosyl radical (Y(*)) cofactor that initiates nucleotide reduction. Pre-steady-state experiments with wild type R1 or C754S/C759S-R1 and R2 were carried out to determine which step(s) are rate-limiting and whether both active sites of R1 can catalyze nucleotide reduction. Rapid chemical quench experiments monitoring dCDP formation gave k(obs) of 9 +/- 4 s(-1) with an amplitude of 1.7 +/- 0.4 equiv. This amplitude, generated in experiments with pre-reduced R1 (3 or 15 microM) in the absence of reductant, indicates that both monomers of R1 are active. Stopped-flow UV-vis spectroscopy monitoring the concentration of the Y(*) failed to reveal any changes from 2 ms to seconds under similar conditions. These pre-steady-state experiments, in conjunction with the steady-state turnover numbers for dCDP formation of 2-14 s(-1) at RNR concentrations of 0.05-0.4 microM (typical assay conditions), reveal that the rate-determining step is a physical step prior to rapid nucleotide reduction and rapid tyrosine reoxidation to Y(*). Steady-state experiments conducted at RNR concentrations of 3 and 15 microM, typical of pre-steady-state conditions, suggest that, in addition to the slow conformational change(s) prior to chemistry, re-reduction of the active site disulfide to dithiol or a conformational change accompanying this process can also be rate-limiting.  相似文献   

13.
Kinetic analysis of respiratory nitrate reductase from Escherichia coli K12   总被引:13,自引:0,他引:13  
Purified respiratory nitrate reductase from Escherichia coli is able to use either reduced viologen dyes or quinols as the electron donor and nitrate, chlorate, or bromate as the electron acceptor. When reduced viologen dyes act as the electron donor, the enzyme follows a compulsory-order, "Theorell-Chance" mechanism, in which it is an enzyme-nitrate complex that is reduced rather than the free enzyme. In contrast, if quinols are used as the electron donor, then the enzyme operates by a two-site, enzyme-substitution mechanism. Partial proteolysis of the cytochrome b containing holoenzyme by trypsin results in loss of cytochrome b and in cleavage of one of the enzyme's subunits. The cytochrome-free derivative exhibits a viologen dye dependent activity that is indistinguishable from that of the holoenzyme, but it is incapable of catalyzing the quinol-dependent reaction. The quinol-dependent, but not the viologen dye dependent, activity is inhibited irreversibly by exposure to diethyl pyrocarbonate and reversibly by treatment with 2-n-heptyl-4-hydroxyquinoline N-oxide. We conclude that the holoenzyme has two independent and spatially distinct active sites, one for quinol oxidation and the other for nitrate reduction.  相似文献   

14.
Knapp KG  Swartz JR 《FEBS letters》2004,559(1-3):66-70
Escherichia coli thioredoxin reductase (TR) and glutathione reductase (GR) are dimeric proteins that require a flavin adenine dinucleotide (FAD) cofactor for activity. A cell-free protein synthesis (CFPS) reaction supplemented with FAD was used to produce TR at 760 microg/ml with 89% of the protein being soluble. GR accumulated to 521 microg/ml in a cell-free reaction with 71% solubility. The TR produced was fully active with a specific activity of 1390 min(-1). The GR had a specific activity of 139 U/mg, which is significantly more active than reported for GR purified from cells. The specific activity for both TR and GR decreased without FAD supplementation. This research demonstrates that CFPS can be used to produce enzymes that are multimeric and require a cofactor.  相似文献   

15.
The kinetics of the NADP+- and phosphate-dependent oxidation of glutamic acid 5-semialdehyde are consistent with a rapid-equilibrium random order mechanism. The Km for DL-pyrroline-5-carboxylic acid is 2.5 mM, for NADP+ is 0.05 mM and for phosphate is 0.35 mM. The Vmax is approx. 8.0 units per mg protein. The reaction is highly specific for the DL-pyrroline-5-carboxylic acid and NADP+, but a number of divalent anions can substitute for phosphate. NADPH is competitive with respect to all three substrates and an analog of gamma-glutamyl phosphate, 3-(phosphonoacetylamido)-L-alanine, is competitive with respect to DL-pyrroline-5-carboxylic acid and non-competitive with respect to NADP+ and phosphate, suggesting dead-end complex formation.  相似文献   

16.
17.
18.
Kinetic analysis and mechanistic aspects of autoxidation of catechins   总被引:3,自引:0,他引:3  
A peroxidase-based bioelectrochemical sensor of hydrogen peroxide (H(2)O(2)) and a Clark-type oxygen electrode were applied to continuous monitoring and kinetic analysis of the autoxidation of catechins. Four major catechins in green tea, (-)-epicatechin, (-)-epicatechin gallate, (-)-epigallocatechin, and (-)-epigallocatechin gallate, were used as model compounds. It was found that dioxygen (O(2)) is quantitatively reduced to H(2)O(2). The initial rate of autoxidation is suppressed by superoxide dismutase and H(+), but is independent of buffer capacity. Based on these results, a mechanism of autoxidation is proposed; the initial step is the one-electron oxidation of the B ring of catechins by O(2) to generate a superoxide anion (O(2)(*-)) and a semiquinone radical, as supported in part by electron spin resonance measurements. O(2)(*-) works as a stronger one-electron oxidant than O(2) against catechins and is reduced to H(2)O(2). The semiquinone radical is more susceptible to oxidation with O(2) than fully reduced catechins. The autoxidation rate increases with pH. This behavior can be interpreted in terms of the increase in the stability of O(2)(*-) and the semiquinone radical with increasing pH, rather than the acid dissociation of phenolic groups. Cupric ion enhances autoxidation; most probably it functions as a catalyst of the initial oxidation step of catechins. The product cuprous ion can trigger a Fenton reaction to generate hydroxyl radical. On the other hand, borate ion suppresses autoxidation drastically, due to the strong complex formation with catechins. The biological significance of autoxidation and its effectors are also discussed.  相似文献   

19.
20.
Strict and facultative anaerobes depend on a class III ribonucleotide reductase for their growth. These enzymes are the sole cellular catalysts for de novo biosynthesis of the deoxyribonucleotides needed for DNA chain elongation and repair. In its active form, the class III ribonucleotide reductase from Escherichia coli contains a free radical located on the G681 residue which is essential for the activation of the ribonucleotide substrate toward its reduction. The 3D structure of the homologous enzyme from bacteriophage T4 has revealed the presence of a metal center bound to four conserved cysteine residues. In this report we identify the metal of the E. coli enzyme as Zn. We show that the presence of Zn in this site protects the protein from proteolysis and prevents the formation of disulfide bridges within it. Finally, we show with the fully Zn-loaded reductase that thioredoxin or small thiols are dispensable for the formation of the glycyl radical. However, they are necessary for obtaining high turnover numbers, suggesting that they intervene in radical transfer steps subsequent to the formation of the glycyl radical.  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