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 906 毫秒
1.
Tyr114 and Tyr197 are highly conserved residues in the active site of human glutathione reductase, Tyr114 in the glutathione disulfide (GSSG) binding site and Tyr197 in the NADPH site. Mutation of either residue has profound effects on catalysis. Y197S and Y114L have 17% and 14% the activity of the wild-type enzyme, respectively. Mutation of Tyr197, in the NADPH site, leads to a decrease in Km for GSSG, and mutation of Tyr114, in the GSSG site, leads to a decrease in Km for NADPH. This behavior is predicted for enzymes operating by a ping-pong mechanism where both half-reactions partially limit turnover. Titration of the wild-type enzyme or Y114L with NADPH proceeds in two phases, Eox to EH2 and EH2 to EH2-NADPH. In contrast, Y197S reacts monophasically, showing that excess NADPH fails to enhance the absorbance of the thiolate-FAD charge-transfer complex, the predominant EH2 form of glutathione reductase. The reductive half-reactions of the wild-type enzyme and of Y114L are similar; FAD reduction is fast (approximately 500 s-1 at 4 degreesC) and thiolate-FAD charge-transfer complex formation has a rate of 100 s-1. In Y197S, these rates are only 78 and 5 s-1, respectively. The oxidative half-reaction, the rate of reoxidation of EH2 by GSSG, of the wild-type enzyme is approximately 4-fold faster than that of Y114L. These results are consistent with Tyr197 serving as a gate in the binding of NADPH, and they indicate that Tyr114 assists the acid catalyst His467'.  相似文献   

2.
Tyrosine 34 is a prominent and conserved residue in the active site of the manganese superoxide dismutases in organisms from bacteria to man. We have prepared the mutant containing the replacement Tyr 34 --> Phe (Y34F) in human manganese superoxide dismutase (hMnSOD) and crystallized it in two different crystal forms, orthorhombic and hexagonal. Crystal structures of hMnSOD Y34F have been solved to 1.9 A resolution in a hexagonal crystal form, denoted as Y34Fhex, and to 2.2 A resolution in an orthorhombic crystal form, denoted as Y34Fortho. Both crystal forms give structures that are closely superimposable with that of wild-type hMnSOD, with the phenyl rings of Tyr 34 in the wild type and Phe 34 in the mutant very similar in orientation. Therefore, in Y34F, a hydrogen-bonded relay that links the metal-bound hydroxyl to ordered solvent (Mn-OH to Gln 143 to Tyr 34 to H2O to His 30) is broken. Surprisingly, the loss of the Tyr 34 hydrogen bonds resulted in large increases in stability (measured by Tm), suggesting that the Tyr 34 hydroxyl does not play a role in stabilizing active-site architecture. The functional role of the side chain hydroxyl of Tyr 34 can be evaluated by comparison of the Y34F mutant with the wild-type hMnSOD. Both wild-type and Y34F had kcat/Km near 10(9) M-1 s-1, close to diffusion-controlled; however, Y34F showed kcat for maximal catalysis smaller by 10-fold than the wild type. In addition, the mutant Y34F was more susceptible to product inhibition by peroxide than the wild-type enzyme. This activity profile and the breaking of the hydrogen-bonding chain at the active site caused by the replacement Tyr 34 --> Phe suggest that Tyr 34 is a proton donor for O2* - reduction to H2O2 or is involved indirectly by orienting solvent or other residues for proton transfer. Up to 100 mM buffers in solution failed to enhance catalysis by either Y34F or the wild-type hMnSOD, suggesting that protonation from solution cannot enhance the release of the inhibiting bound peroxide ion, likely reflecting the enclosure of the active site by conserved residues as shown by the X-ray structures. The increased thermostability of the mutant Y34F and equal diffusion-controlled activity of Y34F and wild-type enzymes with normal superoxide levels suggest that evolutionary conservation of active-site residues in metalloenzymes reflects constraints from extreme rather than average cellular conditions. This new hypothesis that extreme rather than normal substrate concentrations are a powerful constraint on residue conservation may apply most strongly to enzyme defenses where the ability to meet extreme conditions directly affects cell survival.  相似文献   

3.
The high-affinity streptavidin-biotin complex is characterized by an extensive hydrogen-bonding network. A study of hydrogen-bonding energetics at the ureido oxygen of biotin has been conducted with site-directed mutations at Asn 23, Ser 27, and Tyr 43. A new competitive biotin binding assay was developed to provide direct equilibrium measurements of the alterations in Kd. S27A, Y43F, Y43A, N23A, and N23E mutants display DeltaDeltaG degrees at 37 degrees C relative to wild-type streptavidin of 2.9, 1.2, 2.6, 3.5, and 2.6 kcal/mol, respectively. The equilibrium-binding enthalpies for all of the mutants were measured by isothermal titration calorimetry, and the Y43A and N23A mutants display large decreases in the equilibrium binding enthalpy at 25 degrees C of 8.9 and 6.9 kcal/mol, respectively. The S27A and N23E mutants displayed small decreases in binding enthalpy of 1.6 and 0.9 kcal/mol relative to wild-type, while the Y43F mutant displayed a -2.6 kcal/mol increase in the binding enthalpy at 25 degrees C. At 37 degrees C, the Y43A and N23A mutants display decreases of 7.8 and 7.9 kcal/mol, respectively, while the S27A, N23E, and Y43F mutants displayed decreases of 4.9, 3.7, and 1.2 kcal/mol relative to wild-type. Kinetic analyses were also conducted to probe the contributions of the hydrogen bonds to the activation barrier. Wild-type streptavidin at 37 degrees C displays a koff of (4.1 +/- 0.3) x 10(-5) s-1, and the conservative Y43F, S27A, and N23A mutants displayed increases in koff to (20 +/- 1) x 10(-5) s-1, (660 +/- 40) x 10(-5) s-1, and (1030 +/- 220) x 10(-)5 s-1, respectively. The Y43A and N23E mutants displayed 93-fold and 188-fold increases in koff, respectively. Activation energies and enthalpies for each of the mutants were determined by transition-state analysis of the dissociation rate temperature dependence. All of the mutants except Y43F display large reductions in the activation enthalpy. The Y43F mutant has a more positive activation enthalpy, and thus a more favorable activation entropy that underlies the overall reduction in the activation barrier. For the most conservative mutant at each ureido oxygen hydrogen-bonding position, bound-state alterations account for most of the energetic changes in a single transition-state model, suggesting that the ureido oxygen hydrogen-bonding interactions are broken in the dissociation transition state.  相似文献   

4.
A cephalosporin acylase from Pseudomonas strain N176 hydrolyses both 7-beta-(4-carboxybutanamido)-cephalosporanic acid (glutarylcephalosporanic acid) and cephalosporin C to 7-amino-cephalosporanic acid. However, its productivity in the original host was low and its activity against cephalosporin C was not sufficient for direct large-scale production of 7-amino-cephalosporanic acid. In order to overcome these problems, we established a high-level expression system for the acylase in Escherichia coli. Tyr270 in the acylase is reported to play an important role in the interaction with glutarylcephalosporanic acid, as determined from the reaction with an affinity-label reagent, 7 beta-(6-bromohexanoylamido) cephalosporanic acid [Ishii, Y., Saito, Y., Sasaki, H., Uchiyama, F., Hayashi, M., Nakamura, S. & Niwa, M. (1994) J. Ferment. Bioeng. 77, 598-603] and modification with tetranitromethane [Nobbs, T. J., Ishii, Y., Fujimura, T., Saito, Y. & Niwa, M. (1994) J. Ferment. Bioeng. 77, 604-609]. From carbamoylation with potassium cyanate and site-directed point mutagenesis of the cephalosporin C acylase, we have deduced that Tyr270 exists at a position where it can interact with a residue (possibly Ser239) corresponding to inactivation by carbamoylation. We mutated Met269 and Ala271 of the acylase and found that mutation of Met269 to Tyr or Phe caused a 1.6-fold and 1.7-fold increase, respectively, of specific activity against cephalosporin C as compared to that of the wild-type enzyme. Kinetic studies of these mutants revealed that their kcat values increased, although their Km values against cephalosporin C were not changed. These data indicate that the mutation of Met269 near Tyr270 induces a minor conformational change to increase the stability of the activated complex with the enzyme and cephalosporin C. In particular, a mutant in which Met269 was replaced by Tyr was 2.5-fold more efficient in converting cephalosporin C to 7-amino-cephalosporanic acid than the wild-type enzyme under conditions similar to those in a bio-reactor system.  相似文献   

5.
Two catalytically inactive mutant forms of 1-aminocyclopropane-1-carboxylate (ACC) synthase, Y85A and K273A, were mixed in low concentrations of guanidine hydrochloride (GdnHCl). About 15% of the wild-type activity was recovered (theoretical 25% for a binomial distribution), proving that the functional unit of the enzyme is a dimer, or theoretically, a higher order oligomer. The enzyme catalyzes the conversion of S-adenosyl-L-methionine (SAM) to ACC. The value of kcat/KM is 1.2 x 10(6) M-1 s-1 at pH 8.3. Viscosity variation experiments with glycerol and sucrose as viscosogenic reagents showed that this reaction is nearly 100% diffusion controlled. The sensitivity to viscosity for the corresponding reaction of the less reactive Y233F mutant is much reduced, thus the latter reaction serves as a control for that of the wild-type enzyme. The kcat/KM vs pH profile for wild-type enzyme exhibits pKa values of 7.5 and 8.9. The former is assigned to the pKa of the alpha-amino group of SAM, while the latter corresponds to the independently determined spectrophotometric pKa of the internal aldimine. The kcat vs pH profile exhibits similar pKas, which means that the above pKa values are not perturbed in the Michaelis complex. The phenolic hydroxyl group of Tyr233 forms a hydrogen bond to the 3'-O- of PLP. The spectral and kinetic pKa (kcat/KM) values of the Y233F mutant are not identical (spectral 10.2, kinetic 8.7). A model that accounts quantitatively for these data posits two parallel pathways to the external aldimine for this mutant, the minor one has the alpha-amino group free base form of SAM reacting with the protonated imine form of the enzyme with kcat/KM approximately 6.0 x 10(3) M-1 s-1, while the major pathway involves reaction of the aldehyde form of PLP with SAM with kcat/KM approximately 7.0 x 10(5) M-1 s-1. The spectral pKa is defined only by the less reactive species.  相似文献   

6.
3Beta-hydroxysteroid dehydrogenase/steroid delta5-isomerase (3beta-HSD/isomerase) was expressed by baculovirus in Spodoptera fungiperda (Sf9) insect cells from cDNA sequences encoding the human wild-type I (placental) enzyme and the human type I mutant- Y253F. The wild-type and Y253F enzymes were each purified as a single, homogeneous protein from a suspension of the Sf9 cells. Ultraviolet (UV) spectral analyses showed that the wild-type enzyme induced changes in the UV spectrum of the competitive isomerase inhibitor, 19-nortestosterone, and the Y253F mutant did not. The wild-type isomerase required activation by coenzyme to produce the spectral shift. Activation of isomerase by NADH produced a greater change in the 19-nortestosterone spectrum than activation by NAD+. These observations provide direct evidence that Tyr253 functions as the general acid (proton donor) in the isomerase reaction mechanism. Furthermore, the coenzyme-activation profiles support our proposed two-step enzyme mechanism in which NADH produced by the 3beta-HSD activity induces the enzyme to assume the isomerase conformation.  相似文献   

7.
NADH oxidase from Amphibacillus xylanus is a potent alkyl hydroperoxide reductase in the presence of the small disulfide-containing protein (AhpC) of Salmonella typhimurium. In the presence of saturating AhpC, kcat values for reduction of hydroperoxides are approximately 180 s-1, and the double mutant flavoprotein enzyme C337S/C340S cannot support hydroperoxide reduction (Niimura, Y., Poole, L. B., and Massey, V. (1995) J. Biol. Chem. 270, 25645-25650). Kinetics of reduction of wild-type and mutant enzymes are reported here with wild-type enzyme; reduction by NADH was triphasic, with consumption of 2.6 equivalents of NADH, consistent with the known composition of one FAD and two disulfides per subunit. Rate constants for the first two phases (each approximately 200 s-1) where FAD and one disulfide are reduced are slightly greater than kcat values for AhpC-linked hydroperoxide reduction. The rate constant for the third phase (reduction to the 6-electron level) is too small for catalysis. Only the first phase of the wild-type enzyme occurs with the mutant enzyme. These results and the stoichiometry of NADH consumption indicate Cys337 and Cys340 as the active site disulfide of the flavoprotein and that electrons from FADH2 must pass through this disulfide to reduce the disulfide of AhpC.  相似文献   

8.
Site-directed mutagenesis was used to identify key amino acid residues of the cholesterol oxidase from Streptomyces sp., which catalyzes the oxidation of cholesterol and the isomerization of 5-cholesten-3-one. Eight mutant enzymes were constructed and the following amino acid substitutions were identified: N318A, N318H, E356A, E356D, H441A, H441N, N480A and N480Q. Mutants N318A and N318H retained both oxidation and isomerization activities. The mutant E356D retained oxidation but not isomerization activity. On the other hand, mutants N480A and N480Q showed no oxidation activity but retained their isomerization activities. The two catalytic reactions, oxidation and isomerization, in cholesterol oxidase were thus successfully separated. When the H441A or H441N mutation was introduced, both the oxidase and isomerase activities were completely lost. The H441, E356 and N480 residues thus appear to participate in the catalysis of cholesterol oxidase, whereas N318 does not. An analysis of the products of these mutant enzymes suggested that the previously proposed 6-hydroxylation reaction by cholesterol oxidase is actually autooxidation from 5-cholesten-3-one. Kinetic studies of the purified wild-type and mutant enzymes showed that the k(cat)/Km values for oxidation in E356D and for isomerization in N480A increased six- and threefold, respectively, over those in the wild-type. These mutational effects and the reaction mechanisms are discussed in terms of the three-dimensional structure of the enzyme constructed on the basis of homology modeling.  相似文献   

9.
The azomethine (Schiff base) linkage between the epsilon-amino group of active-site lysine 258 and the carbonyl moiety of enzyme-bound pyridoxal 5'-phosphate (PLP) normally exhibits absorbance maxima at ca. 360 (high-pH form) or ca. 430 nm (low-pH form). However, the absorbance maximum is shifted from 358 to 386 nm, a value which is similar to that of free PLP (lambda max = 388 nm), in a mutant form of Escherichia coli aspartate aminotransferase (AATase) in which tyrosine 225, which normally donates a hydrogen bond to the phenolate function of PLP, has been replaced with phenylalanine (Y225F). This spectral shift suggested that PLP binds to Y225F as the free aldehyde. The following evidence from isotope-edited classical Raman spectroscopy proves conclusively that the near-UV spectrum is anomalous and that PLP is bound to Y225F as a Schiff base: (1) A strong cofactor peak at 1630 cm-1 in the holoenzyme-minus-apoenzyme difference spectrum of the unprotonated form of Y225F is red-shifted by 18 cm-1 in enzyme labeled with 15N at lysine 258 and other positions. (2) This isotope-induced red shift is similar to that observed in the unprotonated form of the model Schiff base, PLP-valine. (3) The Raman spectrum of Y225F is unchanged in H(2)18O, while peaks at ca. 1670 cm-1 in the spectrum of free PLP or in that of a mutant of AATase in which Lys-258 is replaced with Ala, are red-shifted by ca. 30 cm-1 in H(2)18O.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
Dihydropyrimidine dehydrogenase catalyzes, in the rate-limiting step of the pyrimidine degradation pathway, the NADPH-dependent reduction of uracil and thymine to dihydrouracil and dihydrothymine, respectively. The porcine enzyme is a homodimeric iron-sulfur flavoprotein (2 x 111 kDa). C671, the residue postulated to be in the uracil binding site and to act as the catalytically essential acidic residue of the enzyme oxidative half-reaction, was replaced by an alanyl residue. The mutant enzyme was overproduced in Escherichia coli DH5alpha cells, purified to homogeneity, and characterized in comparison with the wild-type species. An extinction coefficient of 74 mM-1 cm-1 was determined at 450 nm for the wild-type and mutant enzymes. Chemical analyses of the flavin, iron, and acid-labile sulfur content of the enzyme subunits revealed similar stoichiometries for wild-type and C671A dihydropyrimidine dehydrogenases. One FAD and one FMN per enzyme subunit were found. Approximately 16 iron atoms and 16 acid-labile sulfur atoms were found per wild-type and mutant enzyme subunit. The C671A dihydropyrimidine dehydrogenase mutant exhibited approximately 1% of the activity of the wild-type enzyme, thus preventing its steady-state kinetic analysis. Therefore, the ability of the C671A mutant and, for comparison, of the wild-type enzyme species to interact with reaction substrates, products, or their analogues were studied by absorption spectroscopy. Both enzyme forms did not react with sulfite. The wild-type and mutant enzymes were very similar to each other with respect to the spectral changes induced by binding of the reaction product NADP+ or of its nonreducible analogue 3-aminopyridine dinucleotide phosphate. Uracil also induced qualitatively and quantitatively similar absorbance changes in the visible region of the absorbance spectrum of the two enzyme forms. However, the calculated Kd of the enzyme-uracil complex was significantly higher for the C671A mutant (9.1 +/- 0.7 microM) than for the wild-type dihydropyrimidine dehydrogenase (0.7 +/- 0.09 microM). In line with these observations, the two enzyme forms behaved in a similar way when titrated anaerobically with a NADPH solution. Addition of an up to 10-fold excess of NADPH to both dihydropyrimidine dehydrogenase forms led to absorbance changes consistent with reduction of approximately 0.5 flavin per subunit, with no indication of reduction of the enzyme iron-sulfur clusters. Absorbance changes consistent with reduction of both enzyme flavins were obtained by removing NADP+ with a NADPH-regenerating system. On the contrary, the two enzyme species differed significantly with respect to their reactivity with dihydrouracil. Addition of dihydrouracil to the wild-type enzyme species, under anaerobic conditions, led to absorbance changes that could be interpreted to result from both partial flavin reduction and the formation of a complex between the enzyme and (dihydro)uracil. In contrast, only spectral changes consistent with formation of a complex between the oxidized enzyme and dihydrouracil were observed when a C671A mutant enzyme solution was titrated with this compound. Furthermore, enzyme-monitored turnover experiments were carried out anaerobically in the presence of a limiting amount of NADPH and excess uracil with the two enzyme forms in a stopped-flow apparatus. These experiments directly demonstrated that the substitution of an alanyl residue for C671 in dihydropyrimidine dehydrogenase specifically prevents enzyme-catalyzed reduction of uracil. Finally, sequence analysis of dihydropyrimidine dehydrogenase revealed that it exhibits a modular structure; the N-terminal region, similar to the beta subunit of bacterial glutamate synthases, is proposed to be responsible for NADPH binding and oxidation with reduction of the FAD cofactor of dihydropyrimidine dehydrogenase. The central region, similar to the FMN subunit of dihydroorotate dehydrogenases, is likely to harbor the site o  相似文献   

11.
The conserved Trp residue within helix 5 of the N-lobe of human serum transferrin (hTF/2N, 40 kDa) has been mutated to Tyr. NMR and CD spectra and energy calculations show that the mutation causes little perturbation of the overall structure of hTF/2N although the chelating agent Tiron removed Fe3+ from the mutant protein about three times faster than from wild-type hTF/2N. 1H-NMR resonances of residues in the Leu122-Trp128-Ile132 hydrophobic patch are assigned both by ring current calculations and with the aid of the mutation. [1H, 15N]-NMR resonances for 11 of the 14 Tyr residues were observed in the spectra of 15N-Tyr-hTF/2N and a resonance for Tyr128 was assignable in spectra of the mutant. The 15N resonance of Y128 was sensitive to oxalate and Ga3+ binding, and Ga3+ binding perturbed 15N resonances for most of the Tyr residues. Since these are well distributed over the N-lobe, it can be concluded that metal-induced structural changes are not merely local to the binding site.  相似文献   

12.
The kinetics of intramolecular electron transfer between flavin and heme in Saccharomyces cerevisiae flavocytochrome b2 were investigated by performing potentiometric titrations and temperature-jump experiments on the recombinant wild type and Y143F and Y254F mutants. The midpoint potential of heme was determined by monitoring redox titrations spectrophotometrically, and that of semiquinone flavin/reduced flavin (Fsq/Fred) and oxidized flavin (Fox)/Fsq couples by electron paramagnetic resonance experiments at room temperature. The effects of pyruvate on the kinetic and thermodynamic parameters were also investigated. At room temperature, pH 7.0 and I = 0.1 M, the redox potential of the Fsq/Fred, Fox/Fsq, and oxidized heme/reduced heme (Hox/Hred) couples were -135, -45, and -3 mV, respectively, in the wild-type form. Although neither the mutations nor excess pyruvate did appreciably modify the potential of the heme or that of the Fsq/Fred couple, they led to variable positive shifts in the potential of the Fox/Fsq couple, thus modulating the driving force that characterizes the reduction of heme by the semiquinone in the -42 to +88 mV range. The relaxation rates measured at 16 degreesC in temperature-jump experiments were independent of the protein concentrations, with absorbance changes corresponding to the reduction of the heme. Two relaxation processes were clearly resolved in wild-type flavocytochrome b2 (1/tau1 = 1500 s-1, 1/tau2 = 200 +/- 50 s-1) and were assigned to the reactions whereby the heme is reduced by Fred and Fsq, respectively. The rate of the latter reaction was determined in the whole series of proteins. Its variation as a function of the driving force is well described by the expression obtained from electron-transfer theories, which provides evidence that the intramolecular electron transfer is not controlled by the dynamics of the protein.  相似文献   

13.
The recently reported X-ray structures of cytochrome oxidase reveal structures that are likely proton-conducting channels. One of these channels, leading from the negative aqueous surface to the heme a3/CuB bimetallic center, contains a lysine as a central element. Previous work has shown that this lysine (K362 in the oxidase from Rhodobacter sphaeroides) is essential for cytochrome c oxidase activity. The data presented demonstrate that the K362M mutant is impeded in the reduction of the heme a3/CuB bimetallic center, probably by interfering with the intramolecular movement of protons. The reduction of the heme-copper center is required prior to the reaction with dioxygen to form the so-called peroxy intermediate (compound P). This block can be by-passed to some extent by the addition of H2O2, which can react with the enzyme without prereduction of the heme-copper center and can then be reduced to water using electrons from cytochrome c. Hence, the K362M mutant, though lacking oxidase activity, exhibits cytochrome c peroxidase activity. Rapid mixing techniques have been used to determine the kinetics of this peroxidase activity at concentrations of H2O2 up to 0.5 M. The Km for peroxide is about 50 mM and the Vmax is 50 electrons s-1, which is considerably slower than the turnover that can be obtained for the oxidase activity of the wild-type enzyme (1200 s-1). The turnover of the mutant oxidase with H2O2 appears to be limited by the rate of reaction of the enzyme with peroxide to form compound P, rather than the rate of reduction of compound P to water by cytochrome c. The data require a reexamination of the proposed roles of the putative proton-conducting channels.  相似文献   

14.
Flavocytochrome b2 from Saccharomyces cerevisiae couples L-lactate dehydrogenation to cytochrome c reduction in the mitochondrial intermembrane space. The catalytic cycle for this process can be described in terms of five consecutive electron-transfer events. L-Lactate dehydrogenation results in the two-electron reduction of FMN. The two electrons are individually passed to b2-heme (intramolecular electron transfer) and then onto cytochrome c (intermolecular electron transfer). At 25 degrees C, I 0.10, in the presence of saturating concentrations of ferricytochrome c and L-lactate, the catalytic cycle progresses with rate constant 104 (+/- 5) s-1 [per L-lactate oxidized; Miles, C. S., Rouviere-Fourmy, N., Lederer, F., Mathews, F. S., Reid, G. A., & Chapman, S. K. (1992) Biochem. J. 285, 187-192]. Stopped-flow spectrophotometry has been used to show that the major rate-limiting step in the catalytic cycle is electron transfer from flavin semiquinone to b2-heme. This conclusion is based on the observation that pre-steady-state flavin oxidation by ferricytochrome c takes place at 120 s-1. Although flavin oxidation involves several other electron transfer steps, these are considered too fast to contribute significantly to the rate constant. It was also shown that the reaction product, pyruvate, is able to inhibit pre-steady-state flavin oxidation (Ki = 40 +/- 17 mM) consistent with reports that it acts as a noncompetitive inhibitor in the steady state at high concentrations [Ki = 30 mM; Lederer, F. (1978) Eur. J. Biochem, 88, 425-431]. This novel way of measuring the electron transfer rate constant is directly applicable to the catalytic cycle and has enabled us to derive a self-consistent model for it, based also on data collected for enzyme reduction [Miles, C. S., Rouviere-Fourmy, N., Lederer, F., Mathews, F. S., Reid, G. A., & Chapman, S. K. (1992) Biochem. J. 285, 187-192] and its interaction with cytochrome c [Daff, S., Sharp, R. E., Short, D. M., Bell, C., White, P., Manson, F. D. C., Reid, G. A., & Chapman, S. K. (1996) Biochemistry 35, 6351-6357]. Rapid-freezing quenched-flow EPR has been used to confirm the model by demonstrating that during steady-state turnover of the enzyme approximately 75% of the flavin is in the semiquinone oxidation state.  相似文献   

15.
The flavin-containing enzyme dihydroorotate dehydrogenase (DHOD) catalyzes the oxidation of dihydroorotate (DHO) to orotate, the first aromatic intermediate in pyrimidine biosynthesis. The first structure of a DHOD, the A form of the enzyme from Lactococcus lactis, has recently become known, and some conserved residues were suggested to have a role in the active site [Rowland et al. (1997) Structure 2, 239-252]. In particular, Cys 130 was hypothesized to work as a base, which activates dihydroorotate (DHO) for hydride transfer. By chemical modification and site-directed mutagenesis we have obtained results consistent with this proposal. Cys 130 was susceptible to alkylating reagents, and mutants of Cys 130 (C130A and C130S) showed hardly detectable enzyme activity at pH 8.0, while at pH 10 the C130S mutant enzyme had approximately 1% of wild-type activity. Mutants of Lys 43, Asn 132, and Lys 164 were also constructed. Exchange of Lys 43 to Ala or Glu (K43A and K43E) and of Asn 132 to Ala (N132A) affected both catalysis and substrate binding. Expressed as kcat/KM for DHO, the deterioration of these three mutant enzymes was 10(3)-10(4)-fold. Flavin spectra of the mutant enzymes were not, like the wild-type enzyme, bleached by DHO in stopped-flow experiments, showing that they were deficient with respect to the first half-reaction, namely reduction of FMN by DHO, which was not rate limiting for the wild-type enzyme. The binding interaction between flavin and the reaction product, orotate, could be monitored by a red shift of the flavin absorbance in the wild-type enzyme. The C130A, C130S, and N132A mutant enzymes displayed similar capacity to bind orotate. In contrast, orotate did not change the absorption spectra of the K43 mutant enzymes, although it did inhibit their activity. All of the mutant enzymes, except K164A, contained normal levels of flavin. The results are discussed in relation to the structures of DHODA and other flavoenzymes. The possible acid-base chemistry of Cys 130 is compared to previous work on mammalian dihydropyrimidine dehydrogenases, flavoenzymes, which catalyze the reversed reaction, namely the reduction of pyrimidine bases.  相似文献   

16.
The carbohydrate-binding cleft of Bacillus licheniformis 1,3-1, 4-beta-D-glucan 4-glucanohydrolase is partially covered by the surface loop between residues 51 and 67, which is linked to beta-strand-(87-95) of the minor beta-sheet III of the protein core by a single disulfide bond at Cys61-Cys90. An alanine scanning mutagenesis approach has been applied to analyze the role of loop residues from Asp51 to Arg64 in substrate binding and stability by means of equilibrium urea denaturation, enzyme thermotolerance, and kinetics. The DeltaDeltaGU between oxidized and reduced forms is approximately constant for all mutants, with a contribution of 5.3 +/- 0.2 kcal.mol-1 for the disulfide bridge to protein stability. A good correlation is observed between DeltaGU values by reversible unfolding and enzyme thermotolerance. The N57A mutant, however, is more thermotolerant than the wild-type enzyme, whereas it is slightly less stable to reversible urea denaturation. Mutants with a <2-fold increase in Km correspond to mutations at residues not involved in substrate binding, for which the reduction in catalytic efficiency (kcat/Km) is proportional to the loss of stability relative to the wild-type enzyme. Y53A, N55A, F59A, and W63A, on the other hand, show a pronounced effect on catalytic efficiency, with Km > 2-fold and kcat < 5% of the wild-type values. These mutated residues are directly involved in substrate binding or in hydrophobic packing of the loop. Interestingly, the mutation M58A yields an enzyme that is more active than the wild-type enzyme (7-fold increase in kcat), but it is slightly less stable.  相似文献   

17.
It is thought that only a limited number of residues in the extended binding pocket of coagulation proteases are critical for substrate and inhibitor specificity. A candidate residue from the crystal structures of thrombin and factor Xa (FXa) that may be critical for specificity at the S2 subsite is residue 99. Residue 99 is Tyr in FXa and Thr in activated protein C (APC). To determine the role of residue 99 in S2 specificity, a Gla-domainless mutant of protein C (GDPC) was prepared in which Thr99 was replaced with Tyr of FXa. GDPC T99Y bound Ca2+ and was activated by the thrombin-thrombomodulin complex normally. The T99Y mutant, similar to FXa, hydrolyzed the chromogenic substrates with a Gly at the P2 positions. This mutant was also inhibited by antithrombin (AT) (k2 = 4.2 +/- 0.2 x 10(1) M-1 s-1), and heparin accelerated the reaction >350-fold (k2 = 1.5 +/- 0.1 x 10(4) M-1 s-1). The T99Y mutant, however, did not activate prothrombin but inactivated factor Va approximately 2-fold better than wild type. To try to switch the specificity of FXa, both Tyr99 and Gln192 of FXa were replaced with those of APC in the Gla-domainless factor X (GDFX Y99T/Q192E). This mutant was folded correctly as it bound Ca2+ with a similar affinity as GDFX and was also activated by the Russell's viper venom at similar rate, but it cleaved the chromogenic substrates with a Gly at the P2 positions poorly. The mutant, instead, cleaved the APC-specific chromogenic substrates efficiently. The Y99T/Q192E mutant became resistant to inhibition by AT in the absence of heparin but was inhibited by AT almost normally in the presence of heparin (k2 = 3.4 +/- 0.5 x 10(5) M-1 s-1). The Y99T/Q192E mutant did not inactivate factor Va, and prothrombin activation by this mutant was impaired. These results indicate that 1) residue 99 is critical for enzyme specificity at the S2 subsite, 2) a role for heparin in acceleration of FXa inhibition by AT may involve the S2-P2 modulation, and 3) the exchange of residues 99 and 192 in FXa and APC may switch the enzyme specificity with the chromogenic substrates and inhibitors but not with the natural substrates.  相似文献   

18.
Nitric oxide synthase (NOS) catalyzes the oxidation of L-arginine to citrulline and nitric oxide. C415H and C415A mutants of the neuronal isoform of NOS (nNOS) were expressed in a baculovirus system and purified to homogeneity for spectral analysis and activity measurements. UV-visible spectra of each mutant lacked an observable Soret peak, suggesting that neither mutant contained heme. When reduced in the presence of CO, however, a small Soret centered at 417 nm could be detected for the C415H mutant, further supporting the assignment of C415 as the axial ligand to the heme. In addition to a deficiency in bound heme, neither mutant had any detectable bound tetrahydrobiopterin, as compared to wild-type enzyme, which had a ratio of 0.84 mol of bound pteridine:1 mol of nNOS 160 kDa subunit. The C415H mutant contained bound FAD and FMN at levels of 1.0 +/- 0.1 and 0.9 +/- 0.1 mol/mol of nNOS subunit, respectively. UV-visible spectra of both nNOS mutants retained the distinctive absorbance due to tightly associated oxidized flavin prosthetic groups. Further, the spectra suggested the presence of a neutral flavin semiquinone. Ferricyanide oxidation of the C415A mutant yielded a spectrum that was essentially that of oxidized flavin. Ferricyanide titration showed that the C415A mutant contained approximately 1 reducing equiv. Circular dichroism spectra suggested that each mutant was folded properly, in that both spectra were found to be essentially identical to the spectrum of wild-type nNOS. Neither mutant could synthesize nitric oxide, and neither mutant had the ability to oxidize NADPH unless an exogenous electron acceptor was added. The rate of cytochrome c reduction by each mutant was found to be slightly less, but very similar to the rate (approximately 20 mumol mg-1 min-1) observed with wild-type nNOS. In all cases, the rate of cytochrome c reduction increased approximately 15-fold with the addition of calmodulin. Overall, these spectral and activity data suggest that C415 is the axial heme ligand and that a point mutation at C415 prevents binding of heme and tetrahydrobiopterin without interfering with the global folding or the reductase function of nNOS.  相似文献   

19.
Xanthine dehydrogenase, a molybdenum, iron-sulfur flavoenzyme encoded in the fruit fly Drosophila melanogaster by the rosy gene, has been characterised both from the wild-type and mutant files. Enzyme assays, using a variety of different oxidising and reducing substrates were supplemented by limited molecular characterisation. Four rosy strains showed no detectable activity in any enzyme assay tried, whereas from four wild-type and three rosy mutant strains, those for the [E89K], [L127F] and [L157P]xanthine dehydrogenases (in all of which the mutation is in the iron-sulfur domain), the enzyme molecules, although present at different levels, had extremely similar or identical properties. This was confirmed by purification of one wild-type and one mutant enzyme. [E89K]xanthine dehydrogenase. These both had ultraviolet-visible absorption spectra similar to milk xanthine oxidase. Both were found to be quite stable molecules, showing very high catalytic-centre activities and with little tendency to become degraded by proteolysis or modified by conversion to oxidase or desulfo forms. In three further rosy strains, giving [G353D]xanthine dehydrogenase and [S357F]xanthine dehydrogenase mutated in the flavin domain, and [G1011E]xanthine dehydrogenase mutated in the molybdenum domain, enzyme activities were selectively diminished in certain assays. For the G353D and S357F mutant enzymes activities to NAD+ as oxidising substrate were diminished, to zero for the latter. In addition for [G353D]xanthine dehydrogenase, there was an increase in apparent Km values both for NAD+ and NADH. These findings indicate involvement of this part of the sequence in the NAD(+)-binding site. The G1011E mutation has a profound effect on the enzyme. As isolated and as present in crude extracts of the files, this xanthine dehydrogenase variant lacks activity to xanthine or pterin as reducing substrate, indicating an impairment of the functioning of its molybdenum centre. However, it retains full activity to NADH with dyes as oxidising substrate. Mild oxidation of the enzyme converts it, apparently irreversibly, to a form showing full activity to xanthine and pterin. The nature of the group that is oxidised is discussed in the light of redox potential data. It is proposed that the process involves oxidation of the pterin of the molybdenum cofactor from the tetrahydro to a dihydro oxidation state. This conclusion is fully consistent with recent information [Rom?o, M. J., Archer, M., Moura, I., Moura. J.J.G., LeGall, J., Engh, R., Schneider, M., Hof, P. & Huber, R. (1995) Science 270. 1170-1176) from X-ray crystallography on the structure of a closely related enzyme from Desulfovibrio gigas. It is proposed, that apparent irreversibility of the oxidative activating process for [G1011E]xanthine dehydrogenase, is due to conversion of its pterin to the tricyclic derivative detected by these workers. The data thus provide the strongest evidence available, that the oxidation state of the pterin can have a controlling influence on the activity of a molybdenum cofactor enzyme. Implications regarding pterin incorporation into xanthine dehydrogenase and in relation to other molybdenum enzymes are discussed.  相似文献   

20.
Galactose-1-phosphate uridylyltransferase catalyzes the reaction of UDP-glucose with galactose 1-phosphate (Gal-1-P) to form UDP-galactose and glucose 1-phosphate (Glc-1-P) through a double displacement mechanism, with the intermediate formation of a covalent uridylyl-enzyme (UMP enzyme). Gln 168 in E. coli uridylyltransferase engages in hydrogen bonding with the phosphoryl oxygens of the UMP moiety, which is bonded to His 166 in the intermediate [Wedekind, J. E., Frey, P. A., and Rayment, I. (1996) Biochemistry 35, 11560-11569]. In humans, the point variant Q188R accounts for 60% of galactosemia cases. The corresponding E. coli variant Q168R has been overexpressed and purified. In preparation for kinetic correlation of Q168R and wild-type uridylyltransferases, we tested the kinetic competence of the wild-type UMP-enzyme. At 4 degreesC, the first-order rate constant for uridylylation by UDP-glucose is 281 +/- 18 s-1, and for deuridylylation it is 226 +/- 10 s-1 with Glc-1-P and 166 +/- 10 s-1 with Gal-1-P. Inasmuch as the overall turnover number at 4 degreesC is 62 s-1, the covalent intermediate is kinetically competent. The variant Q168R is uridylylated by UDP-glucose to the extent of about 65% of the potential active sites. Uridylylation reactions of Q168R with UDP-glucose proceed with maximum first-order rate constants of 2.2 x 10(-)4 s-1 and 4.2 x 10(-)4 s-1 at 4 and 27 degreesC, respectively. In experiments with uridylyl-Q168R and glucose-1-P, the mutant enzyme undergoes deuridylylation with maximum first-order rate constants of 4.8 x 10(-)4 s-1 and 1.68 x 10(-)3 s-1 at 4 and 27 degreesC, respectively. The value of Km for uridylylation of Q168R is slightly higher than for the wild-type enzyme, and for deuridylylation it is similar to the wild-type value. The wild-type enzyme undergoes uridylylation and deuridylyation about 10(6) times faster than Q168R. The wild-type activity in the overall reaction is 1.8 x 10(6) times that of Q168R. The wild-type enzyme contains 1.9 mol of Zn+Fe per mole of subunits, whereas the Q168R-variant contains 1.36 mol of Zn+Fe per mole of subunits. The mutation stabilizes the uridylyl-enzyme by 1.2 kcal mol-1 in comparison to the wild-type enzyme. These results show that the low activity of Q168R is not due to overstabilization of the intermediate or to the absence of structural metal ions. Instead, the main defect is very slow uridylylation and deuridylation.  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