共查询到20条相似文献,搜索用时 15 毫秒
1.
Rapid-scan stopped-flow studies of the pH dependence of the reaction between mercuric reductase and NADPH 总被引:1,自引:0,他引:1
The reaction of NADPH with the flavoenzyme mercuric reductase has been studied by rapid-scan stopped-flow spectrophotometry at 5 degrees C in the pH range 5.1-9.5. An intermediate formed within the dead time of the apparatus, and proposed to be an NADPH complex of oxidized enzyme, has an almost pH-independent spectrum. At pH 5.1 the formation of this species is followed by a rapid bleaching (k = 145 s-1) of the main flavin absorption band at 455 nm concomitantly with an absorbance increase around 395 nm. This process, which has a kinetic hydrogen isotope effect of 2.4, becomes less prominent at higher pH values and is not detectable above pH 7. It is suggested that this process includes the formation of a covalent thiol-flavin C-4a derivative stabilized by protonation of the active site. In the presence of an excess of NADPH, the final product of the reaction is probably an NADPH complex of two-electron-reduced enzyme, but below pH 6 the final spectrum becomes less intense suggesting a partial formation of four-electron-reduced enzyme. The spectral changes observed above pH 7 are nearly independent of pH. The first measurable step (k = 48 s-1 at pH 9.5) is thought to include the formation of an NADP+ complex of two-electron-reduced enzyme, while the final step (k = 6.3 s-1 at pH 9.5) results in the above-mentioned NADPH complex with two-electron-reduced enzyme. A minimal kinetic scheme rationalizing the observed pH dependence of the reaction and the observed isotope effects is presented. 相似文献
2.
The reaction of the FAD-containing enzyme, mercuric reductase, with NADPH has been studied by stopped-flow kinetic methods at 25 degrees C, pH 7.3. The results suggest that the reaction involves at least three steps. The first step is very rapid and is essentially complete within the dead time of the stopped-flow apparatus. This step is associated with decreasing absorbances at 340 nm (NADPH) and 455 nm (FAD), whereas there is little formation of the absorbance at 530 nm characterizing 2-electron-reduced enzyme subunits (EH2). The second step involves an increase of the absorbance at 530 nm. The third step results in an increase of the intensity of the long-wavelength band and a change of its shape. A second equivalent of NADPH per FAD is required for this step. It is proposed that the product is an EH2-NADPH complex. In addition to these rapid steps, slow absorbance changes are also observed. 相似文献
3.
M J Moore M D Distefano C T Walsh N Schiering E F Pai 《The Journal of biological chemistry》1989,264(24):14386-14388
The flavoenzyme mercuric ion reductase from Bacillus sp. strain RC607 was purified by dye-ligand affinity chromatography. The protein was crystallized from solutions of high ionic strength, and one of the two crystal forms obtained has proven suitable for x-ray diffraction studies. Preliminary analysis showed that these crystals belong to the tetragonal space group 1422. The unit cell dimensions are a = b = 180.7 A; c = 127.9 A. The diffraction pattern extends to better than 3 A resolution. Crystal density measurements are consistent with one enzyme dimer of 2 x 69,000 Da comprising the asymmetric unit. Trypsin treatment of the native enzyme resulted in the removal of 157 amino acids at the N terminus. After purification, the remaining fragment (amino acids 158-631), which is still fully active in vitro, could be crystallized under the same conditions as native enzyme. Twinning problems, however, did not allow complete analysis of these crystals. 相似文献
4.
Benzoyl-CoA reductase (BCR) catalyzes the ATP-driven transport of two electrons from a reduced 2[4Fe-4S] ferredoxin to the aromatic ring of benzoyl-CoA. A mechanism involving radical species and very low potential electrons similar to the Birch reduction of aromatics has been suggested for this reaction. The redox centers of BCR have previously been identified, by EPR- and M?ssbauer spectroscopy, to be three cysteine-ligated [4Fe-4S] clusters [Boll et al. (2000) J. Biol. Chem. 275, 31857-31868] with redox potentials more negative than -500 mV. In this work, the catalytic cycle of BCR was studied by freeze-quench experiments; the dithionite reduced enzyme was rapidly mixed with equimolar amounts of benzoyl-CoA and excess MgATP plus dithionite, and subjected to EPR spectroscopic analysis. The turnover period of the enzyme under the conditions used was 3 s. The total S = (1)/(2) spin concentration increased 3-fold very rapidly (within approximately 25 ms). In the course of a single turnover the extent of enzyme reduction decreased again, finally reaching the starting value. An increased magnetic interaction of [4Fe-4S] clusters and the rise of an S = (7)/(2) high-spin EPR signal occurred as second simultaneous and transient events (at approximately 200 ms). Previous work showed that binding of the nucleotide affects the magnetic interaction of [4Fe-4S] clusters, whereas hydrolysis of MgATP is required for the switch to high-spin EPR signals. Finally, two novel transient EPR signals with an isotropic line-shape developed maximally in the late phase of the catalytic cycle ( approximately 1-2 s). These signals differed from those of typical free radicals by shifted g values at g = 2.015 and g = 2.033 and by an unusually fast relaxation rate, suggesting an interaction of these paramagnetic species with [4Fe-4S](+1) clusters. On the basis of these results, we present a proposal for a catalytic cycle involving radical species. 相似文献
5.
We have developed and optimized a stopped-flow fluorescence assay for use in studying DNA unwinding catalyzed by Escherichia coli RecBCD helicase. This assay monitors changes in fluorescence resonance energy transfer (FRET) between a pair of fluorescent probes (Cy3 donor and Cy5 acceptor) placed on opposite sides of a nick in duplex DNA. As such, this is an "all-or-none" DNA unwinding assay. Single turnover DNA unwinding experiments were performed using a series of eight fluorescent DNA substrates containing duplex DNA regions ranging from 24 bp to 60 bp. The time-courses obtained by monitoring Cy3 fluorescence display a distinct lag phase that increases with increasing duplex DNA length, reflecting the transient formation of partially unwound DNA intermediates. These Cy3 FRET time-courses are identical with those obtained using a chemical quenched-flow kinetic assay developed previously. The signal from the Cy5 fluorescence probe shows additional effects that appear to specifically monitor the RecD helicase subunit. The continuous nature of this fluorescence assay enabled us to acquire more precise time-courses for many more duplex DNA lengths in a significantly reduced amount of time, compared to quenched-flow methods. Global analysis of the Cy3 and Cy5 FRET time-courses, using an n-step sequential DNA unwinding model, indicates that RecBCD unwinds duplex DNA with an average unwinding rate constant of kU = 200(+/-40) steps s(-1) (mkU = 680(+/-12)bp s(-1)) and an average kinetic step size, m = 3.4 (+/-0.6) bp step(-1) (5 mM ATP, 10 mM MgCl(2), 30 mM NaCl, pH 7.0, 5% (v/v) glycerol, 25.0 degrees C), in excellent agreement with the kinetic parameters determined using quenched-flow techniques. Under these same conditions, the RecBC enzyme unwinds DNA with a very similar rate. These methods will facilitate detailed studies of the mechanisms of DNA unwinding and translocation of the RecBCD and RecBC helicases. 相似文献
6.
Preincubation of the oxidized form of the flavoenzyme mercuric reductase with the reducing substrate, NADPH, or with a high concentration of cysteine (30 mM) results in a substantial increase of the catalytic activity as measured in a standard spectrophotometric assay. Also NADH has some activating effect but NADP+ or EDTA have no effect. In the presence of 1 mM cysteine only one equivalent of NADPH per FAD seems to be required for full activation which occurs after an incubation time of about 10 min. Activated mercuric reductase appears to be stable under anaerobic conditions but eventually returns to the original level of activity in the presence of oxygen. The activated state seems to be stabilized by 1 mM cysteine. Activation of mercuric reductase does not seem to be correlated with a change in the number of reactive thiol groups. The chemical nature of the activation process is not yet understood. Stopped-flow studies have shown that the nonactivated enzyme is practically inactive prior to contact with the substrates. The enzyme is gradually activated during the assay. The kinetics of activation of the 'native' enzyme is biphasic but 'clipped' enzyme, lacking an 85-residue N-terminal domain, is activated in a single first-order process. The progress curves obtained with preactivated enzyme are approximately exponential even at saturating concentrations of NADPH (Km = 0.4 microM at 25 degrees C, pH 7.3) and Hg2+ (Km = 3.2 microM in the presence of 1 mM cysteine). The initial rates yield kcat values of about 13 s-1 per FAD molecule (25 degrees C, pH 7.3). We find no evidence for a thiol-dependent change from a rapid to a slow kinetic phase. The shape of the progress curves presumably depends on product inhibition, but NADP+ is not a sufficiently effective inhibitor to explain the effect fully. 相似文献
7.
BVR-B (biliverdin-IXbeta reductase) also known as FR (flavin reductase) is a promiscuous enzyme catalysing the pyridine-nucleotide-dependent reduction of a variety of flavins, biliverdins, PQQ (pyrroloquinoline quinone) and ferric ion. Mechanistically it is a good model for BVR-A (biliverdin-IXalpha reductase), a potential pharmacological target for neonatal jaundice and also a potential target for adjunct therapy to maintain protective levels of biliverdin-IXalpha during organ transplantation. In a commentary on the structure of BVR-B it was noted that one outstanding issue remained: whether the mechanism was a concerted hydride transfer followed by protonation of a pyrrolic anion or protonation of the pyrrole followed by hydride transfer. In the present study we have attempted to address this question using QM/MM (quantum mechanics/molecular mechanics) calculations. QM/MM potential energy surfaces show that the lowest energy pathway proceeds with a positively charged pyrrole intermediate via two transition states. These initial calculations were performed with His(153) as the source of the proton. However site-directed mutagenesis studies with both the H153A and the H153N mutant reveal that His(153) is not required for catalytic activity. We have repeated the calculation with a solvent hydroxonium donor and obtain a similar energy landscape indicating that protonation of the pyrrole is the most likely first step followed by hydride transfer and that the required proton may come from bulk solvent. The implications of the present study for the design of inhibitors of BVR-A are discussed. 相似文献
8.
Urease, a Ni-containing metalloenzyme, features an activity that has profound medical and agricultural implications. The mechanism of this activity, however, has not been as yet thoroughly established. Accordingly, to improve its understanding, in this study we analyzed the steady-state kinetic parameters of the enzyme (jack bean), K (M) and k (cat), measured at different temperatures and pressures. Such an analysis is useful as it provides information on the molecular nature of the intermediate and transition states of the catalytic reaction. We measured the parameters in a noninteracting buffer using a stopped-flow technique in the temperature range 15-35?°C and in the pressure range 5-132?MPa, the pressure-dependent measurements being the first of their kind performed for urease. While temperature enhanced the activity of urease, pressure inhibited the enzyme; the inhibition was biphasic. Analyzing K (M) provided the characteristics of the formation of the ES complex, and analyzing k (cat), the characteristics of the activation of ES. From the temperature-dependent measurements, the energetic parameters were derived, i.e. thermodynamic ΔH (o) and ΔS (o) for ES formation, and kinetic ΔH ( ≠ ) and ΔS ( ≠ ) for ES activation, while from the pressure-dependent measurements, the binding ΔV (b) and activation [Formula: see text] volumes were determined. The thermodynamic and activation parameters obtained are discussed in terms of the current proposals for the mechanism of the urease reaction, and they are found to support the mechanism proposed by Benini et al. (Structure?7:205-216; 1999), in which the Ni-Ni bridging hydroxide-not the terminal hydroxide-is the nucleophile in the catalytic reaction. 相似文献
9.
Purification and characterization of the flavoenzyme glutathione reductase from rat liver. 总被引:19,自引:0,他引:19
Glutathione reductase from rat liver has been purified greater than 5000-fold in a yield of 20%. The molecular weights of the enzyme and its subunits were estimated to be 125,000 and 60,000, respectively, indicating that the native enzyme is a dimer. The enzyme molecular contains 2 FAD molecules, which are reducible by NADPH, GSH or dithioerythritol. The reduced flavin is instantaneously reoxidized by addition of GSSG. The steady state kinetic data are consistent with a branching reaction mechanism previously proposed for glutathione reductase from yeast (MANNERVIK, B. (1973) Biochem. Biophy. Res. Commun. 53, 1151-1158). This mechanism is also favored by the nonlinear inhibition pattern produced by NADP-+. However, at low GSSG concentrations the rate equation can be approximated by that of a simple ping pong mechanism. NADPH and the mixed disulfide of coenzyme A and GSH were about 10% as active as NADPH and GSSG, respectively, whereas some sulfenyl derivatives related to GSSG were less active as substrates. The pH activity profiles of these substrates differed from that of the NADPH-GSSG substrate pair. 相似文献
10.
Nitroalkane oxidase (NAO) from Fusarium oxysporum catalyzes the oxidation of neutral nitroalkanes to the corresponding aldehydes or ketones with the production of H(2)O(2) and nitrite. The flavoenzyme is a new member of the acyl-CoA dehydrogenase (ACAD) family, but it does not react with acyl-CoA substrates. We present the 2.2 A resolution crystal structure of NAO trapped during the turnover of nitroethane as a covalent N5-FAD adduct (ES*). The homotetrameric structure of ES* was solved by MAD phasing with 52 Se-Met sites in an orthorhombic space group. The electron density for the N5-(2-nitrobutyl)-1,5-dihydro-FAD covalent intermediate is clearly resolved. The structure of ES was used to solve the crystal structure of oxidized NAO at 2.07 A resolution. The c axis for the trigonal space group of oxidized NAO is 485 A, and there are six subunits (1(1)/(2) holoenzymes) in the asymmetric unit. Four of the active sites contain spermine (EI), a weak competitive inhibitor, and two do not contain spermine (E(ox)). The active-site structures of E(ox), EI, and ES* reveal a hydrophobic channel that extends from the exterior of the protein and terminates at Asp402 and the N5 position on the re face of the FAD. Thus, Asp402 is in the correct position to serve as the active-site base, where it is proposed to abstract the alpha proton from neutral nitroalkane substrates. The structures for NAO and various members of the ACAD family overlay with root-mean-square deviations between 1.7 and 3.1 A. The homologous region typically spans more than 325 residues and includes Glu376, which is the active-site base in the prototypical member of the ACAD family. However, NAO and the ACADs exhibit differences in hydrogen-bonding patterns between the respective active-site base, substrate molecules, and FAD. These likely differentiate NAO from the homologues and, consequently, are proposed to result in the unique reaction mechanism of NAO. 相似文献
11.
G Lassmann L Thelander A Gr?slund 《Biochemical and biophysical research communications》1992,188(2):879-887
The reaction of the functional tyrosyl radical in protein R2 of ribonucleotide reductase from E. coli and mouse with the enzyme inhibitor hydroxyurea has been studied by EPR stopped-flow techniques at room temperature. The rate of disappearance of the tyrosyl radical in E. coli protein R2 is k2 = 0.43 M-1 s-1 at 25 degrees C. The reaction follows pseudo-first-order kinetics up to 450 mM hydroxyurea indicating that no saturation by hydroxyurea takes place even at this high concentration. Transient nitroxide-like radicals from hydroxyurea have been detected for the first time in the reaction of hydroxyurea with protein R2 from E. coli and mouse, indicating that 1-electron transfer from hydroxyurea to the tyrosyl radical is the dominating mechanism in the inhibitor reaction. The hydroxyurea radicals appear in low steady-state concentrations during 2-3 half-decay times of the tyrosyl radical and disappear thereafter. 相似文献
12.
Typically, simple flavoprotein oxidases couple the oxidation of their substrates with the formation of hydrogen peroxide without release of significant levels of the superoxide ion. However, two evolutionarily related single-domain sulfhydryl oxidases (Erv2p; a yeast endoplasmic reticulum resident protein and augmenter of liver regeneration, ALR, an enzyme predominantly found in the mitochondrial intermembrane) release up to ~30% of the oxygen they reduce as the superoxide ion. Both enzymes oxidize dithiol substrates via a redox-active disulfide adjacent to the flavin cofactor within the helix-rich Erv domain. Subsequent reduction of the flavin is followed by transfer of reducing equivalents to molecular oxygen. Superoxide release was initially detected using tris(3-hydroxypropyl)phosphine (THP) as an alternative reducing substrate to dithiothreitol (DTT). THP, and other phosphines, showed anomalously high turnover numbers with Erv2p and ALR in the oxygen electrode, but oxygen consumption was drastically suppressed upon the addition of superoxide dismutase. The superoxide ion initiates a radical chain reaction promoting the aerobic oxidation of phosphines with the formation of hydrogen peroxide. Use of a known flux of superoxide generated by the xanthine/xanthine oxidase system showed that one superoxide ion stimulates the reduction of 27 and 4.5 molecules of oxygen using THP and tris(2-carboxyethyl)phosphine (TCEP), respectively. This superoxide-dependent amplification of oxygen consumption by phosphines provides a new kinetic method for the detection of superoxide. Superoxide release was also observed by a standard chemiluminescence method using a luciferin analogue (MCLA) when 2 mM DTT was employed as a substrate of Erv2p and ALR. The percentage of superoxide released from Erv2p increased to ~65% when monomeric mutants of the normally homodimeric enzyme were used. In contrast, monomeric multidomain quiescin sulfhydryl oxidase enzymes that also contain an Erv FAD-binding fold release only 1-5% of their total reduced oxygen species as the superoxide ion. Aspects of the mechanism and possible physiological significance of superoxide release from these Erv-domain flavoproteins are discussed. 相似文献
13.
Mercuric ion reductase (MerA) catalyzes the reduction of Hg(II) to Hg(0) as the last step in the bacterial mercury detoxification pathway. A member of the flavin disulfide oxidoreductase family, MerA contains an FAD prosthetic group and redox-active disulfide in its active site. However, the presence of these two moieties is not sufficient for catalytic Hg(II) reduction, as other enzyme family members are potently inhibited by mercurials. We have previously identified a second pair of active site cysteines (Cys558 Cys559 in the Tn501 enzyme) unique to MerA, that are essential for high levels of mercuric ion reductase activity [Moore, M. J., & Walsh, C. T. (1989) Biochemistry 28, 1183; Miller, S. M., et al. (1989) Biochemistry 28, 1194]. In this paper, we have examined the individual roles of Cys558 and Cys559 by site-directed mutagenesis of each to alanine. Phenotypic analysis indicates that both merA mutations result in a total disruption of the Hg(II) detoxification pathway in vivo, while characterization of the purified mutant enzymes in vitro shows each to have differential effects on catalytic function. Compared to wild-type enzyme, the C558A mutant shows a 20-fold reduction in kcat and a 10-fold increase in Km, for an overall decrease in catalytic efficiency of 200-fold in kcat/Km. In contrast, mutation of Cys559 to alanine results in less than a 2-fold reduction in kcat and an increase in Km of only 4-5 fold for an overall decrease in catalytic efficiency of only ca. 10-fold in vitro. From these results, it appears that Cys558 plays a more important role in forming the reducible complex with Hg(II), while both Cys558 and Cys559 seem to be involved in efficient scavenging (i.e., tight binding) of Hg(II). 相似文献
14.
The falvoenzyme L-alpha-hydroxy acid oxidase from rat kidney [T.H Cromartie and C.T. Walsh (1975), Biochemistry 14, 2588] fails to catalyze the elimination of HCl form D,L-beta-chlorolactate, although this compound is a substrate for oxidation by the enzyme. Deuterium isotope effects demonstrate that proton removal from the alpha carbon of alpha-hydroxy acids is fully rate limiting, a finding in agreement with observations on L-lactate dehydrogenase from yeast [F. Lederer (1974), Eur. J. Biochem. 46, 393] which also does not promote elimination from D,L-beta-chlorolactate. Both D-alpha-hydroxy acid oxidase were found to be rapidly and irreversibly inactivated by the acetylenic substrate 1-hydroxy-3-butynoate. The partially purified dehydrogenase was observed to be inactivated within 10 min by 6.8 times 10(-8) M hydroxybutynoate. For the more extensively studied oxidase, inactivation was found to occur after 25 catalytic events, inactivation occurring by covalent addition of the inactivator to the coenzyme. A stoichimometry of one molecule of hydroxybutynoate per flavine was found, and the time course of inactivation was unaffected by the presence of thiols. The oxidase could also be inactivated by prolonged incubation of the enzyme with 2-hydroxy-3-butenoate, and inactivation which could be completely prevented by the presence of thiolds. Since the inactivation with hydroxybutenoate also left the flavine coenzyme unaltered, the inactivation was attributed to Michael addition of nucleophiles on the enzyme of the ketobutenoate product. Several 4-alkyl-substitued 2-hydroxy-3-butynoates were also observed to inactivate the oxidase by both coenzyme modification and random addition to the apoenzyme. It is proposed that the inactivation may occur by nucleophilic addition of a C4 allenic carbanion to the oxidized flavine coenzyme. 相似文献
15.
Nucleotide sequence of the Thiobacillus ferrooxidans chromosomal gene encoding mercuric reductase 总被引:9,自引:0,他引:9
The nucleotide sequence of the Thiobacillus ferrooxidans chromosomal mercuric-reductase-encoding gene (merA) has been determined. The merA gene contains 1635 bp, and shares 78.2% and 76.6% sequence homology with the transposon, Tn501, and plasmid R100 merA genes, respectively. From the sequence, a 545-amino acid (aa) polypeptide was deduced, and comparison with those of Tn501 and R100 revealed 80.6% and 80.0% homology, respectively, at the aa sequence level. Divergence among the three merA aa sequences was clustered within a specific region (aa positions 41-87). By analysis of codon usage frequency, it is speculated that the T. ferrooxidans merA gene originated from Tn501, R100, or a common ancestral gene, but not from T. ferrooxidans itself. 相似文献
16.
S. Ghosh P. C. Sadhukhan J. Chaudhuri D. K. Ghosh A. Mandal 《Journal of applied microbiology》1999,86(1):7-12
Mercury resistance determinants in bacteria are often plasmid-borne or transposon-mediated. Mercuric reductase, one of the proteins encoded by the mercury resistance operon, catalyses a unique reaction in which mercuric ions, Hg (II), are reduced to mercury metal Hg(O) using NADPH as a source of reducing power. Mercuric reductase was purified from Azotobacter chroococcum SS2 using Red A dye matrix affinity chromatography. Freshly purified preparations of the enzyme showed a single band on polyacrylamide gel electrophoresis under non-denaturing conditions. After SDS-polyacrylamide gel electrophoresis of the freshly prepared enzyme, two protein bands, a major and a minor one, were observed with molecular weight 69 000 and 54 000, respectively. The molecular weight of the native enzyme as determined by gel filtration in Sephacryl S-300 was 142 000. The Km of Hg2+ -reductase for HgCl2 was 11·11 μmol l−1 . Titration with 5,5'-dithiobis (2-nitrobenzoate) demonstrated that two enzyme–SH groups become kinetically accessible on reduction with NADPH. 相似文献
17.
J E Gready 《Biochemistry》1985,24(18):4761-4766
Two mechanisms for facilitating hydride ion transfer from NADPH involving preprotonation of the pteridine rings of the dihydrofolate reductase substrates folate and dihydrofolate have been investigated by ab initio quantum mechanical methods. Protonation energies and effective solution pKas have been calculated for four protonated forms, three of which are nonpreferred in aqueous solution and therefore not directly accessible to experimental study. The pattern and degree of redistribution of the positive charge over the component rings of the N-heterobicyclic pi-system in these protonated forms have been analyzed in terms of changes in the electron populations of the ring atoms and total ring charges. The effects of such changes in promoting hydride ion transfer to C7 in folate and C6 in dihydrofolate have been evaluated by considering the extent of development of partial carbonium ion character at these carbon atoms and also the degree of electron deficiency in the pyrazine ring as a whole. The results illustrate that perturbations due, for instance, to protonation may be propagated by pi-electron coupling effects over medium-range distances of 4-6 A across the pteridine ring. The two mechanisms have been assessed in terms of the calculated absolute and relative pKas of the protonated species taking into account experimental information regarding possible stabilization of these forms in the enzyme active site and also the effectiveness of the various protonations in assisting the hydride ion transfer step. Judged against these criteria, the theoretical results favor the generally proposed mechanism involving preprotonation of N8 in folate and N5 in dihydrofolate. However, some support was also found for the alternative novel mechanism involving O4-protonation of both folate and dihydrofolate. 相似文献
18.
Among the three closely related enzymes, lipoamide dehydrogenase, mercuric reductase, and glutathione reductase only the latter is inhibited by 2,4,6-trinitrobenzenesulfonate (TNBS). On the other hand, all three enzymes exhibit high rates of TNBS-dependent NADPH oxidation. In the case of glutathione reductase and mercuric reductase this TNBS-dependent activity displays substrate inhibition by excess of NADPH and is strongly stimulated by NADP+. The stimulation is especially pronounced with mercuric reductase, 25-fold under some conditions. Neither substrate inhibition nor stimulation by NAD+ is observed with lipoamide dehydrogenase. 相似文献
19.
Venkitakrishnan RP Zaborowski E McElheny D Benkovic SJ Dyson HJ Wright PE 《Biochemistry》2004,43(51):16046-16055
Escherichia coli dihydrofolate reductase (DHFR) has several flexible loops surrounding the active site that play a functional role in substrate and cofactor binding and in catalysis. We have used heteronuclear NMR methods to probe the loop conformations in solution in complexes of DHFR formed during the catalytic cycle. To facilitate the NMR analysis, the enzyme was labeled selectively with [(15)N]alanine. The 13 alanine resonances provide a fingerprint of the protein structure and report on the active site loop conformations and binding of substrate, product, and cofactor. Spectra were recorded for binary and ternary complexes of wild-type DHFR bound to the substrate dihydrofolate (DHF), the product tetrahydrofolate (THF), the pseudosubstrate folate, reduced and oxidized NADPH cofactor, and the inactive cofactor analogue 5,6-dihydroNADPH. The data show that DHFR exists in solution in two dominant conformational states, with the active site loops adopting conformations that closely approximate the occluded or closed conformations identified in earlier X-ray crystallographic analyses. A minor population of a third conformer of unknown structure was observed for the apoenzyme and for the disordered binary complex with 5,6-dihydroNADPH. The reactive Michaelis complex, with both DHF and NADPH bound to the enzyme, could not be studied directly but was modeled by the ternary folate:NADP(+) and dihydrofolate:NADP(+) complexes. From the NMR data, we are able to characterize the active site loop conformation and the occupancy of the substrate and cofactor binding sites in all intermediates formed in the extended catalytic cycle. In the dominant kinetic pathway under steady-state conditions, only the holoenzyme (the binary NADPH complex) and the Michaelis complex adopt the closed loop conformation, and all product complexes are occluded. The catalytic cycle thus involves obligatory conformational transitions between the closed and occluded states. Parallel studies on the catalytically impaired G121V mutant DHFR show that formation of the closed state, in which the nicotinamide ring of the cofactor is inserted into the active site, is energetically disfavored. The G121V mutation, at a position distant from the active site, interferes with coupled loop movements and appears to impair catalysis by destabilizing the closed Michaelis complex and introducing an extra step into the kinetic pathway. 相似文献
20.
Emerson JP Coulter ED Phillips RS Kurtz DM 《The Journal of biological chemistry》2003,278(41):39662-39668
The steady state kinetics of a Desulfovibrio (D.) vulgaris superoxide reductase (SOR) turnover cycle, in which superoxide is catalytically reduced to hydrogen peroxide at a [Fe(His)4(Cys)] active site, are reported. A proximal electron donor, rubredoxin, was used to supply reducing equivalents from NADPH via ferredoxin: NADP+ oxidoreductase, and xanthine/xanthine oxidase was used to provide a calibrated flux of superoxide. SOR turnover in this system was well coupled, i.e. approximately 2O*2 reduced:NADPH oxidized over a 10-fold range of superoxide flux. The reduction of the ferric SOR active site by reduced rubredoxin was independently measured to have a second-order rate constant of approximately 1 x 10(6) m-1 s-1. Analysis of the kinetics showed that: (i) 1 microM SOR can convert a 10 microM/min superoxide flux to a steady state superoxide concentration of 10(-10) m, during which SOR turns over about once every 6 s, (ii) the diffusion-controlled reaction of reduced SOR with superoxide is the slowest process during turnover, and (iii) neither ligation nor deligation of the active site carboxylate of SOR limits the turnover rate. An intracellular SOR concentration on the order of 10 microM is estimated to be the minimum required for lowering superoxide to sublethal levels in aerobically growing SOD knockout mutants of Escherichia coli. SORs from Desulfovibrio gigas and Treponema pallidum showed similar turnover rates when substituted for the D. vulgaris SOR, whereas superoxide dismutases showed no SOR activity in our assay. These results provide quantitative support for previous suggestions that, in times of oxidative stress, SORs efficiently divert intracellular reducing equivalents to superoxide. 相似文献