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 62 毫秒
1.
The substrate reactivity of the flavoenzyme 2-hydroxybiphenyl 3-monooxygenase (EC, HbpA) was changed by directed evolution using error-prone PCR. In situ screening of mutant libraries resulted in the identification of proteins with increased activity towards 2-tert-butylphenol and guaiacol (2-methoxyphenol). One enzyme variant contained amino acid substitutions V368A/L417F, which were inserted by two rounds of mutagenesis. The double replacement improved the efficiency of substrate hydroxylation by reducing the uncoupled oxidation of NADH. With guaiacol as substrate, the two substitutions increased V(max) from 0.22 to 0.43 units mg(-1) protein and decreased the K'(m) from 588 to 143 microm, improving k'(cat)/K'(m) by a factor of 8.2. With 2-tert-butylphenol as the substrate, k'(cat) was increased more than 5-fold. Another selected enzyme variant contained amino acid substitution I244V and had a 30% higher specific activity with 2-sec-butylphenol, guaiacol, and the "natural" substrate 2-hydroxybiphenyl. The K'(m) for guaiacol decreased with this mutant, but the K'(m) for 2-hydroxybiphenyl increased. The primary structure of HbpA shares 20.1% sequence identity with phenol 2-monooxygenase from Trichosporon cutaneum. Structure homology modeling with this three-domain enzyme suggests that Ile(244) of HbpA is located in the substrate binding pocket and is involved in accommodating the phenyl substituent of the phenol. In contrast, Val(368) and Leu(417) are not close to the active site and would not have been obvious candidates for modification by rational design.  相似文献   

2.
Synthesis of 3-tert-butylcatechol by an engineered monooxygenase   总被引:1,自引:0,他引:1  
Recombinant Escherichia coli JM101 was used for the in vivo biocatalytic synthesis of 3-tert-butyl- catechol. The bacterial strain synthesized the laboratory-evolved variant HbpA(T2) of 2-hydroxybiphenyl 3-monooxygenase (HbpA, EC 1.14.13.44) from Pseudomonas azelaica HBP1. The mutant enzyme HbpA(T2) is able to hydroxylate 2-tert-butylphenol to the corresponding catechol, a reaction that is not catalyzed by the wild-type enzyme. The biotransformation was performed in a 3-L bioreactor for 24 h. To mitigate the toxicity of the 2-tert-butylphenol starting material, we applied a limited substrate feed. Continuous in situ product removal with the hydrophobic resin Amberlite XAD-4 was used to separate the product from culture broth. In addition, binding to the resin stabilized the product, which was important because 3-tert-butylcatechol is very labile in aqueous solution. The productivity of the process was 63 mg L(-1) h(-1) so that after 24 h, 3.0 g of 3-tert-butylcatechol were isolated. Down-stream processing consisted of two steps. First, bound 2-tert-butylphenol and 3-tert-butylcatechol were eluted from Amberlite XAD-4 with methanol. Second, the two compounds were separated over neutral aluminum oxide, which selectively binds the produced catechol but not the phenol substrate. The final purity of 3-tert-butylcatechol was greater than 98%.  相似文献   

3.
Previous work showed that random mutagenesis produced a mutant of toluene ortho-monooxygenase (TOM) of Burkholderia cepacia G4 containing the V106A substitution in the hydroxylase -subunit (TomA3) that changed the color of the cell suspension from wild-type brown to green in rich medium. Here, DNA shuffling was used to isolate a random TOM mutant that turned blue due to mutation TomA3 A113V. To better understand the TOM reaction mechanism, we studied the specificity of indole hydroxylation using a spectrum of colored TOM mutants expressed in Escherichia coli TG1 and formed as a result of saturation mutagenesis at TomA3 positions A113 and V106. Colonies expressing these altered enzymes ranged in color from blue through green and purple to orange; and the enzyme products were identified using thin-layer chromatography, high performance liquid chromatography, and liquid chromatography–mass spectroscopy. Derived from the single TOM template, enzymes were identified that produced primarily isoindigo (wild-type TOM), indigo (A113V), indirubin (A113I), and isatin (A113H and V106A/A113G). The discovery that wild-type TOM formed isoindigo via C-2 hydroxylation of the indole pyrrole ring makes this the first oxygenase shown to form this compound. Variant TOM A113G was unable to form indigo, indirubin, or isoindigo (did not hydroxylate the indole pyrrole ring), but produced 4-hydroxyindole and unknown yellow compounds from C-4 hydroxylation of the indole benzene ring. Mutations at V106 in addition to A113G restored C-3 indole oxidation, so along with C-2 indole oxidation, isatin, indigo, and indirubin were formed. Other TomA3 V106/A113 mutants with hydrophobic, polar, or charged amino acids in place of the Val and/or Ala residues hydroxylated indole at the C-3 and C-2 positions, forming isatin, indigo, and indirubin in a variety of distributions. Hence, for the first time, a single enzyme was genetically modified to produce a wide range of colors from indole.  相似文献   

4.
Oxygenases are useful catalysts for the selective incorporation of molecular oxygen into hydrocarbons. Here, we report on the application of isolated, cell free 2-hydroxybiphenyl 3-monooxygenase (HbpA) as catalyst for the regio- and chemospecific hydroxylation of 2-hydroxybiphenyl to 2,3-dihydroxybiphenyl. The catalyst was prepared from recombinant Escherichia coli using expanded bed adsorption chromatography and could be stored without significant loss of activity in lyophilized form. The reaction was performed in an aerated and thermostated simple stirred glass vessel in an aqueous (20% v/v)/organic (80% v/v) reaction medium. This allowed in situ product recovery preventing substrate and product inhibition of the catalyst as well as decay of the labile product 2,3-dihydroxybiphenyl. Enzymatic regeneration of reduced nicotinamide cofactors was achieved using the formate/formate dehydrogenase system. We obtained volumetric productivities of up to 0.45 g l−1 h−1. No significant decrease of productivity was observed within 7 h and more. Product purification (purity 92%) was achieved using solid phase extraction with aluminum oxide followed by crystallization as a polishing step (purity>99%).

To our knowledge, these results show for the first time the perspectives of integrated enzyme and cofactor regeneration-based biocatalytic processes in organic/aqueous emulsions, coupled with in situ product recovery.  相似文献   


5.
Asp222 is an invariant residue in all known sequences of aspartate aminotransferases from a variety of sources and is located within a distance of strong ionic interaction with N(1) of the coenzyme, pyridoxal 5'-phosphate (PLP), or pyridoxamine 5'-phosphate (PMP). This residue of Escherichia coli aspartate aminotransferase was replaced by Ala, Asn, or Glu by site-directed mutagenesis. The PLP form of the mutant enzyme D222E showed pH-dependent spectral changes with a pKa value of 6.44 for the protonation of the internal aldimine bond, slightly lower than that (6.7) for the wild-type enzyme. In contrast, the internal aldimine bond in the D222A or D222N enzyme did not titrate over the pH range 5.3-9.5, and a 430-nm band attributed to the protonated aldimine persisted even at high pH. The binding affinity of the D222A and D222N enzymes for PMP decreased by 3 orders of magnitude as compared to that of the wild-type enzyme. Pre-steady-state half-transamination reactions of all the mutant enzymes with substrates exhibited anomalous progress curves comprising multiphasic exponential processes, which were accounted for by postulating several kinetically different enzyme species for both the PLP and PMP forms of each mutant enzyme. While the replacement of Asp222 by Glu yielded fairly active enzyme species, the replacement by Ala and Asn resulted in 8600- and 20,000-fold decreases, respectively, in the catalytic efficiency (kmax/Kd value for the most active species of each mutant enzyme) in the reactions of the PLP form with aspartate. In contrast, the catalytic efficiency of the PMP form of the D222A or D222N enzyme with 2-oxoglutarate was still retained at a level as high as 2-10% of that of the wild-type enzyme. The presteady-state reactions of these two mutant enzymes with [2-2H]aspartate revealed a deuterium isotope effect (kH/kD = 6.0) greater than that [kH/kD = 2.2; Kuramitsu, S., Hiromi, K., Hayashi, H., Morino, Y., & Kagamiyama, H. (1990) Biochemistry 29, 5469-5476] for the wild-type enzyme. These findings indicate that the presence of a negatively charged residue at position 222 is particularly critical for the withdrawal of the alpha-proton of the amino acid substrate and accelerates this rate-determining step by about 5 kcal.mol-1. Thus it is concluded that Asp222 serves as a protein ligand tethering the coenzyme in a productive mode within the active site and stabilizes the protonated N(1) of the coenzyme to strengthen the electron-withdrawing capacity of the coenzyme.  相似文献   

6.
Cytochrome P450 BM-3 from Bacillus megaterium catalyzed NADPH-supported indole hydroxylation under alkaline conditions with homotropic cooperativity toward indole. The activity was also found with the support of H2O2, tert-butyl hydroperoxide (tBuOOH), or cumene hydroperoxide (CuOOH). Enhanced activity and heterotropic cooperativity were observed in CuOOH-supported hydroxylation, and both the Hill coefficient and substrate concentration required for half-maximal activity in the CuOOH-supported reaction were much lower than those in the H2O2-, tBuOOH-, or NADPH-supported reactions. CuOOH greatly enhanced NADPH consumption and indole hydroxylation in the NADPH-supported reaction. However, when CuOOH was replaced by tBuOOH or H2O2, heterotropic cooperativity was not observed. Spectral studies also confirmed that CuOOH stimulated indole binding to P450 BM-3. Interestingly, a mutant enzyme with enhanced indole-hydroxylation activity, F87V (Phe87 was replaced by Val), lost homotropic cooperativity towards indole and heterotropic cooperativity towards CuOOH, indicating that the active-site structure affects the cooperativities.  相似文献   

7.
The kinetics of Coprinus cinereus peroxidase-catalyzed 1-naphthol, 2-naphthol, and 4-hydroxybiphenyl oxidation was investigated. The initial rates of the naphthols' and 4-hydroxybiphenyl oxidations were linearly dependent on enzyme concentration. The rates depended on substrate concentration and saturated at concentrations above 100 microM of hydrogen peroxide, 25-50 microM of naphthols, and 10 microM of 4-hydroxybiphenyl. At the peroxide concentration 100 microM calculated K(m) and the maximal rate (V(max)) were 74.7 microM and 0.53 microM/sec or 175 microM and 2.0 microM/sec for 1- or 2-naphthol, respectively, and 29.68 microM and 0.42 microM/sec for 4-hydroxybiphenyl. Kinetic measurements of exhaustive naphthol and 4-hydroxybiphenyl oxidation showed that peroxidase is inactivated during the oxidation of the substrates. Different factors and additives, water soluble polymers and albumins (PEG, PEI, PL, BSA, HSA), influenced the initial naphthols and 4-hydroxybiphenyl oxidation rates, peroxidase inactivation rates, and the degree of the substrate conversion. Addition of albumin increased turnover number of naphthols oxidation 1.5-4 times. Light scattering increase was observed when peroxidase-catalyzed oxidation reaction was investigated and suggested that insoluble particles were formed during the process. The addition of polymers, change of concentration and ionic strength of the solution as well as the number of other factors influenced the observed light scattering. The number of particles formed during peroxidase-catalyzed naphthols' and 4-hydroxybiphenyl oxidation and their distribution according to size in the interval 2.5-300 microm were detected by particle counting in solutions.  相似文献   

8.
Huang YC  Grodsky NB  Kim TK  Colman RF 《Biochemistry》2004,43(10):2821-2828
Pig heart mitochondrial NADP-dependent isocitrate dehydrogenase requires a divalent metal ion for catalysis, and metal-isocitrate is its preferred substrate. On the basis of the crystal structure of the enzyme-Mn(2+)-isocitrate complex, Asp(252), Asp(275), and Asp(279) were selected as targets for site-directed mutagenesis to evaluate the roles of these residues as ligands of the metal ion. The circular dichroism spectra of the purified mutant enzymes are similar to that of wild-type enzyme indicating there are no appreciable conformational changes. The K(m) values for isocitrate and for Mn(2+) are increased in the asparagine and histidine mutants at positions 252 and 275; while for cysteine mutants at the same positions, the K(m)'s are not changed appreciably. Mutants at position 279 exhibit only a small change in K(m) for isocitrate. These results indicate that Asp(252) and Asp(275) are ligands of enzyme-bound Mn(2+)and influence the binding of Mn(2+)-isocitrate. Cysteine is an acceptable substitute for aspartate as a ligand of Mn(2+). The pK(aes)'s of D252C and D275C enzymes are similar to that of the wild-type enzyme (about 5.2), while the pK(aes) of D279C is a little lower (about 4.7). These findings suggest that the V(max)'s of the D252C, D275C, and D279C enzymes depend on the ionizable form of the same group as in wild-type enzyme and neither Asp(252), Asp(275), nor Asp(279) acts as the general base in the enzymatic reaction. For wild-type enzyme, the pK(aes) varies with the metal ion used with Mg(2+) > Cd(2+) > Mn(2+) > Co(2+), similar to the order of the pK's for these four metal-bound waters. We therefore attribute the pH dependence of V(max) to the deprotonation of the metal-coordinated hydroxyl group of isocitrate bound to isocitrate dehydrogenase.  相似文献   

9.
In an effort to understand the catalytic mechanism of the tryptophan synthase beta-subunit from Salmonella typhimurium, possible functional active site residues have been identified (on the basis of the 3-D crystal structure of the bienzyme complex) and targeted for analysis utilizing site-directed mutagenesis. The chromophoric properties of the pyridoxal 5'-phosphate cofactor provide a particularly convenient and sensitive spectral probe to directly investigate changes in catalytic events which occur upon modification of the beta-subunit. Substitution of Asp for Glu 109 in the beta-subunit was found to alter both the catalytic activity and the substrate specificity of the beta-reaction. Steady-state kinetic data reveal that the beta-reaction catalyzed by the beta E109D alpha 2 beta 2 mutant enzyme complex is reduced 27-fold compared to the wild-type enzyme. Rapid-scanning stopped-flow (RSSF) UV-visible spectroscopy shows that the mutation does not seriously affect the pre-steady-state reaction of the beta E109D mutant with L-serine to form the alpha-aminoacrylate intermediate, E(A-A). Binding of the alpha-subunit specific ligand, alpha-glycerol phosphate (GP) to the alpha 2 beta 2 complex exerts the same allosteric effects on the beta-subunit as observed with the wild-type enzyme. However, the pre-steady-state spectral changes for the reaction of indole with E(A-A) show that the formation of the L-tryptophan quinonoid, E(Q3), is drastically altered. Discrimination against E(Q3) formation is also observed for the binding of L-tryptophan to the mutant alpha 2 beta 2 complex in the reverse reaction. In contrast, substitution of Asp for Glu 109 increases the apparent affinity of the beta E109D alpha-aminoacrylate complex for the indole analogue indoline and results in the increased rate of synthesis of the amino acid product dihydroiso-L-tryptophan. Thus, the mutation affects the covalent bond forming addition reactions and the nucleophile specificity of the beta-reaction catalyzed by the bienzyme complex.  相似文献   

10.
The ability of Methylosinus trichosporium OB3b, expressing soluble methane monooxygenase, to oxidize a range of ortho-substituted biphenyls was examined to better understand how substituents affect both the rate and products of oxidation in comparison to biphenyl. Inhibition of oxidation was observed over the tested substrate range for both biphenyl and ortho-halogenated biphenyls (2-chloro-, 2-bromo-, and 2-iodobiphenyl). No inhibition was observed during the oxidation of 2-hydroxybiphenyl and 2-methylbiphenyl. Analysis of the products of oxidation showed that, depending on the substituent, ring hydroxylation, substituent oxidation, and elimination pathways could occur. The type and abundance of products formed along with the relatively high kinetic isotope effect observed for deuterated vs. nondeuterated biphenyl (k(h)/k(d) = 3.4+/-0.02) are consistent with mechanisms that include both hydrogen abstraction and NIH-shift pathways. Knowledge of these substituent-dependent reaction rates and mechanisms enhances our understanding of the methanotrophic aryl transformation potential and allows for better prediction of the formation of oxidized intermediates by methanotrophic bacteria.  相似文献   

11.
Hydroxylation of biphenyl by the dibenzofuran-degrading yeast Trichosporon mucoides SBUG 801 was studied. Glucose-grown cells degraded 40% of the biphenyl added within the first 24 h of incubation. The first step in the biotransformation pathway was the monohydroxylation of the biaryl compound to produce 2-, 3-, and 4-hydroxybiphenyl. Further oxidation produced seven dihydroxylated intermediates; the second hydroxyl group was added either on the aromatic ring already hydroxylated or on the second ring. Of all metabolites, 2,5-dihydroxybiphenyl accumulated in the supernatant in the highest concentration. The initial hydroxylation favors the 4-position to produce 4-hydroxybiphenyl, which is subsequently hydroxylated to form 3,4-dihydroxybiphenyl. When biphenyl was replaced as a substrate by 4-hydroxybiphenyl, further hydroxylation of the intermediate 3,4-dihydroxybiphenyl resulted in 3,4,4'-trihydroxybiphenyl. Incubation of T. mucoides with biphenyl and 18O2 indicated a monooxygenase-catalyzed reaction in the oxidation of biphenyl. The hydroxylation was inhibited by 1-aminobenzotriazole and metyrapone, known cytochrome P450 inhibitors. These results are very similar to those observed in the biotransformation of biphenyl in mammals.  相似文献   

12.
Earlier, it had been proposed in the laboratories at Halle that a cysteine residue is responsible for the hysteretic substrate activation behavior of yeast pyruvate decarboxylase. More recently, this idea has received support in a series of studies from Rutgers with the identification of residue C221 as the site where substrate is bound to transmit the information to H92, to E91, to W412, and finally to the active center thiamin diphosphate. According to steady-state kinetic assays, the C221A/C222A variant is no longer subject to substrate activation yet is still a well-functioning enzyme. Several further experiments are reported on this variant: (1) The variant exhibits lag phases in the product formation progress curves, which can be attributed to a unimolecular step in the pre-steady-state stage of catalysis. (2) The rate of exchange with solvent deuterium of the thiamin diphosphate C2H atom is slowed by a factor of 2 compared to the wild-type enzyme, suggesting that the reduced activity that results from the substitutions some 20 A from the active center is also seen in the first key step of the reaction. (3) The solvent (deuterium oxide) kinetic isotope effect was found to be inverse on V(max)/K(m) (0.62), and small but normal on V(max) (1.26), virtually ruling out residue C221 as being responsible for the inverse effects reported for the wild-type enzyme at low substrate concentrations. The solvent kinetic isotope effects are compared to those on two related enzymes not subject to substrate activation, Zymomonas mobilis pyruvate decarboxylase and benzoylformate decarboxylase.  相似文献   

13.
Choline oxidase catalyzes the oxidation of choline to glycine betaine, a compatible solute that accumulates in pathogenic bacteria and plants so they can withstand osmotic and temperature stresses. The crystal structure of choline oxidase was determined and refined to a resolution of 1.86 A with data collected at 100 K using synchrotron X-ray radiation. The structure reveals a covalent linkage between His99 Nepsilon2 and FAD C8M atoms, and a 123 A3 solvent-excluded cavity adjacent to the re face of the flavin. A hypothetical model for choline docked into the cavity suggests that several aromatic residues and Glu312 may orient the cationic substrate for efficient catalysis. The role of the negative charge on Glu312 was investigated by engineering variant enzymes in which Glu312 was replaced with alanine, glutamine, or aspartate. The Glu312Ala enzyme was inactive. The Glu312Gln enzyme exhibited a Kd value for choline at least 500 times larger than that of the wild-type enzyme. The Glu312Asp enzyme had a kcat/KO2 value similar to that of the wild-type enzyme but kcat and kcat/Km values that were 230 and 35 times lower, respectively, than in the wild-type enzyme. These data are consistent with the spatial location of the negative charge on residue 312 being important for the oxidation of the alcohol substrate. Solvent viscosity and substrate kinetic isotope effects suggest the presence of an internal equilibrium in the Glu312Asp enzyme prior to the hydride transfer reaction. Altogether, the crystallographic and mechanistic data suggest that Glu312 is important for binding and positioning of the substrate in the active site of choline oxidase.  相似文献   

14.
The usage by enzymes of specific binding pathways for gaseous substrates or products is debated. The crystal structure of the redox enzyme cholesterol oxidase, determined at sub-angstrom resolution, revealed a hydrophobic tunnel that may serve as a binding pathway for oxygen and hydrogen peroxide. This tunnel is formed by a cascade of conformational rearrangements and connects the active site with the exterior surface of the protein. To elucidate the relationship between this tunnel and gas binding and release, three mutant enzymes were constructed to block the tunnel or its putative gate. Mutation of the proposed gating residue Asn485 to Asp or tunnel residue Phe359 or Gly347 to Trp or Asn reduces the catalytic efficiency of oxidation. The K mO 2 increases from 300 +/- 35 microM for the wild-type enzyme to 617 +/- 15 microM for the F359W mutant. The k cat for the F359W mutant-catalyzed reaction decreases 13-fold relative to that of the wild-type-catalyzed reaction. The N485D and G347N mutants could not be saturated with oxygen. Transfer of hydride from the sterol to the flavin prosthetic group is no longer rate-limiting for these tunnel mutants. The steady-state kinetics of both wild-type and tunnel mutant enzymes are consistent with formation of a ternary complex of steroid and oxygen during catalysis. Furthermore, kinetic cooperativity with respect to molecular oxygen is observed with the tunnel mutants, but not with the wild-type enzyme. A rate-limiting conformational change for binding and release of oxygen and hydrogen peroxide, respectively, is consistent with the cooperative kinetics. In the atomic-resolution structure of F359W, the indole ring of the tryptophan completely fills the tunnel and is observed in only a single conformation. The size of the indole is proposed to limit conformational rearrangement of residue 359 that leads to tunnel opening in the wild-type enzyme. Overall, these results substantiate the functional importance of the tunnel for substrate binding and product release.  相似文献   

15.
Because mutations of the ionizable Asp at position 55 of the phosphatidylcholine preferring phospholipase C from Bacillus cereus (PLC(Bc)) to a non-ionizable Asn generate a mutant enzyme (D55N) with 10(4)-fold lower catalytic activity than the wild-type enzyme, we tentatively identified Asp55 as the general base for the enzymatic reaction. To eliminate the alternate possibility that Asp55 is a structurally important amino acid, the X-ray structures of unbound D55N and complexes of D55N with two non-hydrolyzable substrate analogues have been solved and refined to 2.0, 2.0, and 2.3A, respectively. The structures of unbound wild-type PLC(Bc) and a wild-type PLC(Bc)-complex with a non-hydrolyzable substrate analogue do not change significantly as a result of replacing Asp55 with Asn. These observations demonstrate that Asp55 is not critical for the structural integrity of the enzyme and support the hypothesis that Asp55 is the general base in the PLC(Bc)-catalyzed hydrolysis of phospholipids.  相似文献   

16.
Myoglobin (Mb) is used as a model system for other heme proteins and the reactions they catalyze. The latest novel function to be proposed for myoglobin is a P450 type hydroxylation activity of aromatic carbons (Watanabe, Y., and Ueno, T. (2003) Bull. Chem. Soc. Jpn. 76, 1309-1322). Because Mb does not contain a specific substrate binding site for aromatic compounds near the heme, an engineered tryptophan in the heme pocket was used to model P450 hydroxylation of aromatic compounds. The monooxygenation product was not previously isolated because of rapid subsequent oxidation steps (Hara, I., Ueno, T., Ozaki, S., Itoh, S., Lee, K., Ueyama, N., and Watanabe, Y. (2001) J. Biol. Chem. 276, 36067-36070). In this work, a Mb variant (F43W/H64D/V68I) is used to characterize the monooxygenated intermediate. A modified (+16 Da) species forms upon the addition of 1 eq of H2O2. This product was digested with chymotrypsin, and the modified peptide fragments were isolated and characterized as 6-hydroxytryptophan using matrix-assisted laser desorption ionization time-of-flight tandem mass spectroscopy and 1H NMR. This engineered Mb variant represents the first enzyme to preferentially hydroxylate the indole side chain of Trp at the C6 position. Finally, heme extraction was used to demonstrate that both the formation of the 6-hydroxytryptophan intermediate (+16 Da) and subsequent oxidation to form the +30 Da final product are catalyzed by the heme cofactor, most probably via the compound I intermediate. These results provide insight into the mechanism of hydroxylation of aromatic carbons by heme proteins, demonstrating that non-thiolate-ligated heme enzymes can perform this function. This establishes Mb compound I as a model for P450 type aromatic hydroxylation chemistry.  相似文献   

17.
2-Hydroxybiphenyl 3-monooxygenase (EC 1.14.13.44) from Pseudomonas azelaica HBP1 is an FAD-dependent aromatic hydroxylase that catalyzes the conversion of 2-hydroxybiphenyl to 2, 3-dihydroxybiphenyl in the presence of NADH and oxygen. The catalytic mechanism of this three-substrate reaction was investigated at 7 degrees C by stopped-flow absorption spectroscopy. Various individual steps associated with catalysis were readily observed at pH 7.5, the optimum pH for enzyme turnover. Anaerobic reduction of the free enzyme by NADH is a biphasic process, most likely reflecting the presence of two distinct enzyme forms. Binding of 2-hydroxybiphenyl stimulated the rate of enzyme reduction by NADH by 2 orders of magnitude. The anaerobic reduction of the enzyme-substrate complex involved the formation of a transient charge-transfer complex between the reduced flavin and NAD(+). A similar transient intermediate was formed when the enzyme was complexed with the substrate analog 2-sec-butylphenol or with the non-substrate effector 2,3-dihydroxybiphenyl. Excess NAD(+) strongly stabilized the charge-transfer complexes but did not give rise to the appearance of any intermediate during the reduction of uncomplexed enzyme. Free reduced 2-hydroxybiphenyl 3-monooxygenase reacted rapidly with oxygen to form oxidized enzyme with no appearance of intermediates during this reaction. In the presence of 2-hydroxybiphenyl, two consecutive spectral intermediates were observed which were assigned to the flavin C(4a)-hydroperoxide and the flavin C(4a)-hydroxide, respectively. No oxygenated flavin intermediates were observed when the enzyme was in complex with 2, 3-dihydroxybiphenyl. Monovalent anions retarded the dehydration of the flavin C(4a)-hydroxide without stabilization of additional intermediates. The kinetic data for 2-hydroxybiphenyl 3-monooxygenase are consistent with a ternary complex mechanism in which the aromatic substrate has strict control in both the reductive and oxidative half-reaction in a way that reactions leading to substrate hydroxylation are favored over those leading to the futile formation of hydrogen peroxide. NAD(+) release from the reduced enzyme-substrate complex is the slowest step in catalysis.  相似文献   

18.
Moran GR  Phillips RS  Fitzpatrick PF 《Biochemistry》1999,38(49):16283-16289
Tryptophan hydroxylase is a pterin-dependent amino acid hydroxylase that catalyzes the incorporation of one atom of molecular oxygen into tryptophan to form 5-hydroxytryptophan. The substrate specificity and hydroxylation regiospecificity of tryptophan hydroxylase have been investigated using tryptophan analogues that have methyl substituents or nitrogens incorporated into the indole ring. The products of the reactions show that the regiospecificity of tryptophan hydroxylase is stringent. Hydroxylation does not occur at the 4 or 6 carbon in response to changes in substrate topology or atomic charge. 5-Hydroxymethyltryptophan and 5-hydroxy-4-methyltryptophan are the products from 5-methyltryptophan. These products establish that the hydroxylating intermediate is sufficiently potent to hydroxylate benzylic carbons and that the direction of the NIH shift in tryptophan hydroxylase is from carbon 5 to carbon 4. The effects on the V/K values for the amino acids indicate that the enzyme is most sensitive to changes at position 5 of the indole ring. The V(max) values for amino acid hydroxylation differ at most by a factor of 3 from that observed for tryptophan, while the efficiencies of hydroxylation with respect to tetrahydropterin consumption vary 6-fold, consistent with oxygen transfer to the amino acid being partially or fully rate limiting in productive catalysis.  相似文献   

19.
Cytochrome P450 (P450) 2A6 is able to catalyze indole hydroxylation to form the blue dye indigo. The wild-type P450 2A6 enzyme was randomly mutated throughout the whole open reading frame and screened using 4-chloroindole hydroxylation, a substituted indole selected from 30 indole compounds for enhanced color development. Mutants with up to 5-fold increases of catalytic efficiency (k(cat)/K(m)) and 2-fold increases in k(cat) were selected after two rounds of screening. Important residues located both in (e.g., Thr305) and outside the active site (e.g., Ser224) were identified. The study utilized a better substrate for "indigo assay" to obtain new information on the structure-functional relationship of P450 2A6 that was not revealed by previous mutagenesis studies with this enzyme.  相似文献   

20.
Pavon JA  Fitzpatrick PF 《Biochemistry》2006,45(36):11030-11037
Phenylalanine hydroxylase (PheH) and tryptophan hydroxylase (TrpH) catalyze the aromatic hydroxylation of phenylalanine and tryptophan, forming tyrosine and 5-hydroxytryptophan, respectively. The reactions of PheH and TrpH have been investigated with [4-(2)H]-, [3,5-(2)H(2)]-, and (2)H(5)-phenylalanine as substrates. All (D)k(cat) values are normal with Delta117PheH, the catalytic core of rat phenylalanine hydroxylase, ranging from 1.12-1.41. In contrast, for Delta117PheH V379D, a mutant protein in which the stoichiometry between tetrahydropterin oxidation and amino acid hydroxylation is altered, the (D)k(cat) value with [4-(2)H]-phenylalanine is 0.92 but is normal with [3,5-(2)H(2)]-phenylalanine. The ratio of tetrahydropterin oxidation to amino acid hydroxylation for Delta117PheH V379D shows a similar inverse isotope effect with [4-(2)H]-phenylalanine. Intramolecular isotope effects, determined from the deuterium contents of the tyrosine formed from [4-(2)H]-and [3,5(2)H(2)]-phenylalanine, are identical for Delta117PheH and Delta117PheH V379D, suggesting that steps subsequent to oxygen addition are unaffected in the mutant protein. The inverse effects are consistent with the reaction of an activated ferryl-oxo species at the para position of the side chain of the amino acid to form a cationic intermediate. The normal effects on the (D)k(cat) value for the wild-type enzyme are attributed to an isotope effect of 5.1 on the tautomerization of a dienone intermediate to tyrosine with a rate constant 6- to7-fold that for hydroxylation. In addition, there is a slight ( approximately 34%) preference for the loss of the hydrogen originally at C4 of phenylalanine. With (2)H(5)-indole-tryptophan as a substrate for Delta117PheH, the (D)k(cat) value is 0.89, consistent with hydroxylation being rate-limiting in this case. When deuterated phenylalanines are used as substrates for TrpH, the (D)k(cat) values are within error of those for Delta117PheH V379D. Overall, these results are consistent with the aromatic amino acid hydroxylases all sharing the same chemical mechanism, but with the isotope effect for hydroxylation by PheH being masked by tautomerization of an enedione intermediate to tyrosine.  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