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 31 毫秒
1.
Oxygenases are promising biocatalysts for performing selective hydroxylations not accessible by chemical methods. Whereas toluene 4-monooxygenase (T4MO) of Pseudomonas mendocina KR1 hydroxylates monosubstituted benzenes at the para position and toluene ortho-monooxygenase (TOM) of Burkholderia cepacia G4 hydroxylates at the ortho position, toluene 3-monooxygenase (T3MO) of Ralstonia pickettii PKO1 was reported previously to hydroxylate toluene at the meta position, producing primarily m-cresol (R. H. Olsen, J. J. Kukor, and B. Kaphammer, J. Bacteriol. 176:3749-3756, 1994). Using gas chromatography, we have discovered that T3MO hydroxylates monosubstituted benzenes predominantly at the para position. TG1/pBS(Kan)T3MO cells expressing T3MO oxidized toluene at a maximal rate of 11.5 +/- 0.33 nmol/min/mg of protein with an apparent Km value of 250 microM and produced 90% p-cresol and 10% m-cresol. This product mixture was successively transformed to 4-methylcatechol. T4MO, in comparison, produces 97% p-cresol and 3% m-cresol. Pseudomonas aeruginosa PAO1 harboring pRO1966 (the original T3MO-bearing plasmid) also exhibited the same product distribution as that of TG1/pBS(Kan)T3MO. TG1/pBS(Kan)T3MO produced 66% p-nitrophenol and 34% m-nitrophenol from nitrobenzene and 100% p-methoxyphenol from methoxybenzene, as well as 62% 1-naphthol and 38% 2-naphthol from naphthalene; similar results were found with TG1/pBS(Kan)T4MO. Sequencing of the tbu locus from pBS(Kan)T3MO and pRO1966 revealed complete identity between the two, thus eliminating any possible cloning errors. 1H nuclear magnetic resonance analysis confirmed the structural identity of p-cresol in samples containing the product of hydroxylation of toluene by pBS(Kan)T3MO.  相似文献   

2.
Aromatic hydroxylations are important bacterial metabolic processes but are difficult to perform using traditional chemical synthesis, so to use a biological catalyst to convert the priority pollutant benzene into industrially relevant intermediates, benzene oxidation was investigated. It was discovered that toluene 4-monooxygenase (T4MO) of Pseudomonas mendocina KR1, toluene 3-monooxygenase (T3MO) of Ralstonia pickettii PKO1, and toluene ortho-monooxygenase (TOM) of Burkholderia cepacia G4 convert benzene to phenol, catechol, and 1,2,3-trihydroxybenzene by successive hydroxylations. At a concentration of 165 microM and under the control of a constitutive lac promoter, Escherichia coli TG1/pBS(Kan)T4MO expressing T4MO formed phenol from benzene at 19 +/- 1.6 nmol/min/mg of protein, catechol from phenol at 13.6 +/- 0.3 nmol/min/mg of protein, and 1,2,3-trihydroxybenzene from catechol at 2.5 +/- 0.5nmol/min/mg of protein. The catechol and 1,2,3-trihydroxybenzene products were identified by both high-pressure liquid chromatography and mass spectrometry. When analogous plasmid constructs were used, E. coli TG1/pBS(Kan)T3MO expressing T3MO formed phenol, catechol, and 1,2,3-trihydroxybenzene at rates of 3 +/- 1, 3.1 +/- 0.3, and 0.26 +/- 0.09 nmol/min/mg of protein, respectively, and E. coli TG1/pBS(Kan)TOM expressing TOM formed 1,2,3-trihydroxybenzene at a rate of 1.7 +/- 0.3 nmol/min/mg of protein (phenol and catechol formation rates were 0.89 +/- 0.07 and 1.5 +/- 0.3 nmol/min/mg of protein, respectively). Hence, the rates of synthesis of catechol by both T3MO and T4MO and the 1,2,3-trihydroxybenzene formation rate by TOM were found to be comparable to the rates of oxidation of the natural substrate toluene for these enzymes (10.0 +/- 0.8, 4.0 +/- 0.6, and 2.4 +/- 0.3 nmol/min/mg of protein for T4MO, T3MO, and TOM, respectively, at a toluene concentration of 165 microM).  相似文献   

3.
A novel toluene monooxygenase (TMO) six-gene cluster from Bradyrhizobium sp. BTAi1 having an overall 35, 36, and 38 % protein similarity with toluene o-xylene monooxygenase (ToMO) of Pseudomonas sp. OX1, toluene 4-monooxygenase (T4MO) of Pseudomonas mendocina KR1, and toluene-para-monooxygenase (TpMO) of Ralstonia pickettii PKO1, respectively, was cloned and expressed in Escherichia coli TG1, and its potential activity was investigated for aromatic hydroxylation and trichloroethylene (TCE) degradation. The natural substrate toluene was hydroxylated to p-cresol, indicating that the new toluene monooxygenase (T4MO·BTAi1) acts as a para hydroxylating enzyme, similar to T4MO and TpMO. Some shifts in regiospecific hydroxylations were observed compared to the other wild-type TMOs. For example, wild-type T4MO·BTAi1 formed catechol (88 %) and hydroquinone (12 %) from phenol, whereas all the other wild-type TMOs were reported to form only catechol. Furthermore, it was discovered that TG1 cells expressing wild-type T4MO·BTAi1 mineralized TCE at a rate of 0.67?±?0.10 nmol Cl?/h/mg protein. Saturation and site directed mutagenesis were used to generate eight variants of T4MO·BTAi1 at alpha-subunit positions P101, P103, and H214: P101T/P103A, P101S, P101N/P103T, P101V, P103T, P101V/P103T, H214G, and H214G/D278N; by testing the substrates phenol, nitrobenzene, and naphthalene, positions P101 and P103 were found to influence the regiospecific oxidation of aromatics. For example, compared to wild type, variant P103T produced four fold more m-nitrophenol from nitrobenzene as well as produced mainly resorcinol (60 %) from phenol whereas wild-type T4MO·BTAi1 did not. Similarly, variants P101T/P103A and P101S synthesized more 2-naphthol and 2.3-fold and 1.6-fold less 1-naphthol from naphthalene, respectively.  相似文献   

4.
A two-phase system is developed here for converting: (1) benzene to phenol and (2) naphthalene to 2-naphthol, using whole cells expressing wild-type toluene 4-monooxygenase (T4MO) and the alpha subunit variant TmoA I100A from Pseudomonas mendocina KR1. Using the T4MO TmoA I100A variant, the solubility of naphthalene was enhanced and the toxicity of the naphthols was prevented by the use of a water/dioctyl phthalate (80:20, vol%) system which yielded 21-fold more 2-naphthol. More than 99% 2-naphthol was extracted to the dioctyl phthalate phase, dihydroxynaphthalene formation was prevented, 92% 2-naphthol was formed, and 12% naphthalene was converted. Similarly, using 50 vol% dioctyl phthalate, an initial concentration of 3.0 g l−1 (39 mM), and wild-type T4MO, a 51±9% conversion of benzene was obtained and phenol was produced at a purity of 97%. Relative to the one-phase system, there was a 12-fold reduction in the formation of the byproduct catechol.  相似文献   

5.
Directed evolution of toluene ortho-monooxygenase (TOM) of Burkholderia cepacia G4 previously created the hydroxylase alpha-subunit (TomA3) V106A variant (TOM-Green) with increased activity for both trichloroethylene degradation (twofold enhancement) and naphthalene oxidation (six-times-higher activity). In the present study, saturation mutagenesis was performed at position A106 with Escherichia coli TG1/pBS(Kan)TOMV106A to improve TOM activity for both chloroform degradation and naphthalene oxidation. Whole cells expressing the A106E variant had two times better naphthalene-to-1-naphthol activity than the wild-type cells (V(max) of 9.3 versus 4.5 nmol.min(-1).mg of protein(-1) and unchanged K(m)), and the regiospecificity of the A106E variant was unchanged, with 98% 1-naphthol formed, as was confirmed with high-pressure liquid chromatography. The A106E variant degrades its natural substrate toluene 63% faster than wild-type TOM does (2.12 +/- 0.07 versus 1.30 +/- 0.06 nmol.min(-1).mg of protein(-1) [mean +/- standard deviation]) at 91 microM and has a substantial decrease in regiospecificity, since o-cresol (50%), m-cresol (25%), and p-cresol (25%) are formed, in contrast to the 98% o-cresol formed by wild-type TOM. The A106E variant also has an elevated expression level compared to that of wild-type TOM, as evidenced by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Another variant, the A106F variant, has 2.8-times-better chloroform degradation activity based on gas chromatography (V(max) of 2.61 versus 0.95 nmol.min(-1).mg of protein(-1) and unchanged K(m)) and chloride release (0.034 +/- 0.002 versus 0.012 +/- 0.001 nmol.min(-1).mg of protein(-1)). The A106F variant also was expressed at levels similar to those of wild-type TOM and 62%-better toluene oxidation activity than wild-type TOM (2.11 +/- 0.3 versus 1.30 +/- 0.06 nmol.min(-1).mg of protein(-1)). A shift in regiospecificity of toluene hydroxylation was also observed for the A106F variant, with o-cresol (28%), m-cresol (18%), and p-cresol (54%) being formed. Statistical analysis was used to estimate that 292 colonies must be screened for a 99% probability that all 64 codons were sampled during saturation mutagenesis.  相似文献   

6.
Toluene-o-xylene monooxygenase (ToMO) from Pseudomonas stutzeri OX1 has been shown to degrade all chlorinated ethenes individually and as mixtures. Here, DNA shuffling of the alpha hydroxylase fragment of ToMO (TouA) and saturation mutagenesis of the TouA active site residues I100, Q141, T201, F205, and E214 were used to enhance the degradation of chlorinated aliphatics. The ToMO mutants were identified using a chloride ion screen and then were further examined by gas chromatography. Escherichia coli TG1/pBS(Kan)ToMO expressing TouA saturation mutagenesis variant I100Q was identified that has 2.8-fold better trichloroethylene (TCE) degradation activity (apparent V max of 1.77 nmol min−1 mg−1 protein−1 vs 0.63 nmol min−1 mg−1 protein−1). Another variant, E214G/D312N/M399V, has 2.5-fold better cis-1,2-dichloroethylene (cis-DCE) degradation activity (apparent V max of 8.4 nmol min−1 mg−1 protein−1 vs 3.3 nmol min−1 mg−1 protein−1). Additionally, the hydroxylation regiospecificity of o-xylene and naphthalene were altered significantly for ToMO variants A107T/E214A, T201G, and T201S. Variant T201S produced 2.0-fold more 2,3-dimethylphenol (2,3-DMP) from o-xylene than the wild-type ToMO, whereas variant A107T/E214A had 6.0-fold altered regiospecificity for 2,3-DMP formation. Variant A107T/E214A also produced 3.0-fold more 2-naphthol from naphthalene than the wild-type ToMO, whereas the regiospecificity of variant T201S was altered to synthesize 3.0-fold less 2-naphthol, so that it made almost exclusively 1-naphthol (96%). Variant T201G was more regiospecific than variants A107T/E214A and T201S and produced 100% 3,4-DMP from o-xylene and >99% 1-naphthol from naphthalene. Hence, ToMO activity was enhanced for the degradation of TCE and cis-DCE and for the regiospecific hydroxylation of o-xylene and naphthalene through DNA shuffling and saturation mutagenesis.  相似文献   

7.
A primary goal of protein engineering is to control catalytic activity. Here we show that through mutagenesis of three active site residues, the catalytic activity of a multicomponent monooxygenase is altered so that it hydroxylates all three positions of toluene as well as both positions of naphthalene. Hence, for the first time, an enzyme has been engineered so that its regiospecific oxidation of a substrate can be controlled. Through the A107G mutation in the alpha-subunit of toluene para-monooxygenase, a variant was formed that hydroxylated toluene primarily at the ortho-position while converting naphthalene to 1-naphthol. Conversely, the A107T variant produced >98% p-cresol and p-nitrophenol from toluene and nitrobenzene, respectively, as well as produced 2-naphthol from naphthalene. The mutation I100S/G103S produced a toluene para-monooxygenase variant that formed 75% m-cresol from toluene and 100% m-nitrophenol from nitrobenzene; thus, for the first time a true meta-hydroxylating toluene monooxygenase was created.  相似文献   

8.
Of the four investigated Rhodococcus strains (R. rhodochrous 172, R. opacus 4a and 557, and R. rhodnii 135), the first three strains were found to be able to completely transform fluorene when it was present in the medium as the sole source of carbon at a concentration of 12-25 mg/l. At a fluorene concentration of 50-100 mg/l in the medium, the rhodococci transformed 50% of the substrate in 14 days. The addition of casamino acids and sucrose (1-5 g/l) stimulated fluorene transformation, so that R. rhodochrous 172 could completely transform it in 2-5 days. Nine intermediates of fluorene transformation were isolated, purified, and structurally characterized. It was found that R. rhodnii 135 and R. opacus strains 4a and 557 hydroxylated fluorene with the formation of 2-hydroxyfluorene and 2,7-dihydroxyfluorene. R. rhodochrous 172 transformed fluorene via two independent pathways to a greater degree than did the other rhodococci studied.  相似文献   

9.
Toluene-o-xylene monooxygenase (ToMO) from Pseudomonas stutzeri OX1 was found to oxidize nitrobenzene (NB) to form m-nitrophenol (m-NP, 72%) and p-NP (28%) with an initial rate of 0.098 and 0.031 nmol/(min mg protein), respectively. It was also discovered that wild-type ToMO forms 4-nitrocatechol (4-NC) from m-NP and p-NP with an initial rate of 0.15 and 0.0082 nmol/(min mg protein), respectively, and 3-NC (12%) and nitrohydroquinone (NHQ, 88%) from o-NP with an initial rate of 0.11 and 0.8 nmol/(min mg protein), respectively. To increase the oxidation rate and alter the oxidation regiospecificity of nitro aromatics as well as to study the role of the active site residues I100, Q141, T201, and F205 of the alpha hydroxylase fragment of ToMO (TouA), DNA shuffling and saturation mutagenesis were used to generate random mutants. The mutants were initially identified by screening via a rapid agar plate assay and then were further examined by high-performance liquid chromatography (HPLC) and gas chromatography (GC). Several mutants with higher rates of activities and with different regiospecificities were identified; for example, Escherichia coli TG1 cells expressing either TouA mutant M180T/E284G or E214G/D312N/M399V produce 4-NC 4.5- and 20-fold faster than wild-type ToMO (0.037 and 0.16 nmol/min mg protein from p-NP, respectively). TouA mutant A107T/E214A had the regiospecificity of NB changed significantly from 28% to 79% p-NP. From 200 microM NB, TouA variants A101T/M114T, A110T/E392D, M180T/E284G, and E214G/D312N/M399V produce 4-NC whereas wild-type ToMO does not. From m-NP, TouA mutant I100Q produces 4-NC (37%) and NHQ (63%), whereas wild-type ToMO produces only 4-NC (100%). Variant A107T/E214A acts like a para enzyme and forms p-cresol as the major product (93%) from toluene with enhanced activity (2.3-fold), whereas wild-type ToMO forms 32%, 21%, and 47% of o-, m-, and p-cresol, respectively. Hence, the non-specific ToMO was converted into a regiospecific enzyme, which rivals toluene 4-monooxygenase of P. mendocina KR1 and toluene o-monooxygenase of Burkholderia cepacia G4 in its specificity.  相似文献   

10.
Aromatic hydroxylations are important bacterial metabolic processes but are difficult to perform using traditional chemical synthesis, so to use a biological catalyst to convert the priority pollutant benzene into industrially relevant intermediates, benzene oxidation was investigated. It was discovered that toluene 4-monooxygenase (T4MO) of Pseudomonas mendocina KR1, toluene 3-monooxygenase (T3MO) of Ralstonia pickettii PKO1, and toluene ortho-monooxygenase (TOM) of Burkholderia cepacia G4 convert benzene to phenol, catechol, and 1,2,3-trihydroxybenzene by successive hydroxylations. At a concentration of 165 μM and under the control of a constitutive lac promoter, Escherichia coli TG1/pBS(Kan)T4MO expressing T4MO formed phenol from benzene at 19 ± 1.6 nmol/min/mg of protein, catechol from phenol at 13.6 ± 0.3 nmol/min/mg of protein, and 1,2,3-trihydroxybenzene from catechol at 2.5 ± 0.5nmol/min/mg of protein. The catechol and 1,2,3-trihydroxybenzene products were identified by both high-pressure liquid chromatography and mass spectrometry. When analogous plasmid constructs were used, E. coli TG1/pBS(Kan)T3MO expressing T3MO formed phenol, catechol, and 1,2,3-trihydroxybenzene at rates of 3 ± 1, 3.1 ± 0.3, and 0.26 ± 0.09 nmol/min/mg of protein, respectively, and E. coli TG1/pBS(Kan)TOM expressing TOM formed 1,2,3-trihydroxybenzene at a rate of 1.7 ± 0.3 nmol/min/mg of protein (phenol and catechol formation rates were 0.89 ± 0.07 and 1.5 ± 0.3 nmol/min/mg of protein, respectively). Hence, the rates of synthesis of catechol by both T3MO and T4MO and the 1,2,3-trihydroxybenzene formation rate by TOM were found to be comparable to the rates of oxidation of the natural substrate toluene for these enzymes (10.0 ± 0.8, 4.0 ± 0.6, and 2.4 ± 0.3 nmol/min/mg of protein for T4MO, T3MO, and TOM, respectively, at a toluene concentration of 165 μM).  相似文献   

11.
Trichloroethylene (TCE) is the most frequently detected groundwater contaminant, and 1-naphthol is an important chemical manufacturing intermediate. Directed evolution was used to increase the activity of toluene ortho-monooxygenase (TOM) of Burkholderia cepacia G4 for both chlorinated ethenes and naphthalene oxidation. When expressed in Escherichia coli, the variant TOM-Green degraded TCE (2.5 +/- 0.3 versus 1.39 +/- 0.05 nmol/min/mg of protein), 1,1-dichloroethylene, and trans-dichloroethylene more rapidly. Whole cells expressing TOM-Green synthesized 1-naphthol at a rate that was six times faster than that mediated by the wild-type enzyme at a concentration of 0.1 mM (0.19 +/- 0.03 versus 0.029 +/- 0.004 nmol/min/mg of protein), whereas at 5 mM, the mutant enzyme was active (0.07 +/- 0.03 nmol/min/mg of protein) in contrast to the wild-type enzyme, which had no detectable activity. The regiospecificity of TOM-Green was unchanged, with greater than 97% 1-naphthol formed. The beneficial mutation of TOM-Green is the substitution of valine to alanine in position 106 of the alpha-subunit of the hydroxylase, which appears to act as a smaller "gate" to the diiron active center. This hypothesis was supported by the ability of E. coli expressing TOM-Green to oxidize the three-ring compounds, phenanthrene, fluorene, and anthracene faster than the wild-type enzyme. These results show clearly that random, in vitro protein engineering can be used to improve a large multisubunit protein for multiple functions, including environmental restoration and green chemistry.  相似文献   

12.
Wild-type toluene 4-monooxygenase (T4MO) of Pseudomonas mendocina KR1 oxidizes toluene to p-cresol (96%) and oxidizes benzene sequentially to phenol, to catechol, and to 1,2,3-trihydroxybenzene. In this study T4MO was found to oxidize o-cresol to 3-methylcatechol (91%) and methylhydroquinone (9%), to oxidize m-cresol and p-cresol to 4-methylcatechol (100%), and to oxidize o-methoxyphenol to 4-methoxyresorcinol (87%), 3-methoxycatechol (11%), and methoxyhydroquinone (2%). Apparent Vmax values of 6.6 ± 0.9 to 10.7 ± 0.1 nmol/min/ mg of protein were obtained for o-, m-, and p-cresol oxidation by wild-type T4MO, which are comparable to the toluene oxidation rate (15.1 ± 0.8 nmol/min/mg of protein). After these new reactions were discovered, saturation mutagenesis was performed near the diiron catalytic center at positions I100, G103, and A107 of the alpha subunit of the hydroxylase (TmoA) based on directed evolution of the related toluene o-monooxygenase of Burkholderia cepacia G4 (K. A. Canada, S. Iwashita, H. Shim, and T. K. Wood, J. Bacteriol. 184:344-349, 2002) and a previously reported T4MO G103L regiospecific mutant (K. H. Mitchell, J. M. Studts, and B. G. Fox, Biochemistry 41:3176-3188, 2002). By using o-cresol and o-methoxyphenol as model substrates, regiospecific mutants of T4MO were created; for example, TmoA variant G103A/A107S produced 3-methylcatechol (98%) from o-cresol twofold faster and produced 3-methoxycatechol (82%) from 1 mM o-methoxyphenol seven times faster than the wild-type T4MO (1.5 ± 0.2 versus 0.21 ± 0.01 nmol/min/mg of protein). Variant I100L produced 3-methoxycatechol from o-methoxyphenol four times faster than wild-type T4MO, and G103S/A107T produced methylhydroquinone (92%) from o-cresol fourfold faster than wild-type T4MO and there was 10 times more in terms of the percentage of the product. Variant G103S produced 40-fold more methoxyhydroquinone from o-methoxyphenol than the wild-type enzyme produced (80 versus 2%) and produced methylhydroquinone (80%) from o-cresol. Hence, the regiospecific oxidation of o-methoxyphenol and o-cresol was changed for significant synthesis of 3-methoxycatechol, methoxyhydroquinone, 3-methylcatechol, and methylhydroquinone. The enzyme variants also demonstrated altered monohydroxylation regiospecificity for toluene; for example, G103S/A107G formed 82% o-cresol, so saturation mutagenesis converted T4MO into an ortho-hydroxylating enzyme. Furthermore, G103S/A107T formed 100% p-cresol from toluene; hence, a better para-hydroxylating enzyme than wild-type T4MO was formed. Structure homology modeling suggested that hydrogen bonding interactions of the hydroxyl groups of altered residues S103, S107, and T107 influence the regiospecificity of the oxygenase reaction.  相似文献   

13.
Saturation mutagenesis was used to generate 10 mutants of toluene-o-xylene monooxygenase (ToMO) at alpha subunit (TouA) positions A101 and A110: A101G, A101I, A101M, A101VE, A101V, A110G, A110C, A110S, A110P, and A110T; by testing the substrates toluene, o-cresol, m-cresol, p-cresol, phenol, naphthalene, o-methoxyphenol, m-methoxyphenol, p-methoxyphenol, o-xylene, and nitrobenzene, these positions were found to influence the regiospecific oxidation of aromatics. For example, compared to wild-type ToMO, TouA variant A101V produced threefold more 3-methoxycatechol from m-methoxyphenol as well as produced methylhydroquinone from o-cresol whereas wild-type ToMO did not. Similarly, variant A110C synthesized 1.8-fold more o-cresol from toluene and 1.8-fold more 3-methoxycatechol from m-methoxyphenol, and variant A110G synthesized more m-nitrophenol and twofold less p-nitrophenol from nitrobenzene. The A101V and A110C mutations did not affect the rate of reaction with the natural substrate toluene, so the variants had high activity. This is the first report that these or analogous residues influence the catalysis with this class of enzymes. Wild-type ToMO was found to oxidize o-methoxyphenol to methoxyhydroquinone (60%) and 4-methoxyresorcinol (40%), m-methoxyphenol to 4-methoxycatechol (96%) and 3-methoxycatechol (4%), and p-methoxyphenol to 4-methoxycatechol (100%).  相似文献   

14.
Toluene-o-xylene monooxygenase (ToMO) from Pseudomonas stutzeri OX1 oxidizes toluene to 3- and 4-methylcatechol and oxidizes benzene to form phenol; in this study ToMO was found to also form catechol and 1,2,3-trihydroxybenzene (1,2,3-THB) from phenol. To synthesize novel dihydroxy and trihydroxy derivatives of benzene and toluene, DNA shuffling of the alpha-hydroxylase fragment of ToMO (TouA) and saturation mutagenesis of the TouA active site residues I100, Q141, T201, and F205 were used to generate random mutants. The mutants were initially identified by screening with a rapid agar plate assay and then were examined further by high-performance liquid chromatography and gas chromatography. Several regiospecific mutants with high rates of activity were identified; for example, Escherichia coli TG1/pBS(Kan)ToMO expressing the F205G TouA saturation mutagenesis variant formed 4-methylresorcinol (0.78 nmol/min/mg of protein), 3-methylcatechol (0.25 nmol/min/mg of protein), and methylhydroquinone (0.088 nmol/min/mg of protein) from o-cresol, whereas wild-type ToMO formed only 3-methylcatechol (1.1 nmol/min/mg of protein). From o-cresol, the I100Q saturation mutagenesis mutant and the M180T/E284G DNA shuffling mutant formed methylhydroquinone (0.50 and 0.19 nmol/min/mg of protein, respectively) and 3-methylcatechol (0.49 and 1.5 nmol/min/mg of protein, respectively). The F205G mutant formed catechol (0.52 nmol/min/mg of protein), resorcinol (0.090 nmol/min/mg of protein), and hydroquinone (0.070 nmol/min/mg of protein) from phenol, whereas wild-type ToMO formed only catechol (1.5 nmol/min/mg of protein). Both the I100Q mutant and the M180T/E284G mutant formed hydroquinone (1.2 and 0.040 nmol/min/mg of protein, respectively) and catechol (0.28 and 2.0 nmol/min/mg of protein, respectively) from phenol. Dihydroxybenzenes were further oxidized to trihydroxybenzenes with different regiospecificities; for example, the I100Q mutant formed 1,2,4-THB from catechol, whereas wild-type ToMO formed 1,2,3-THB (pyrogallol). Regiospecific oxidation of the natural substrate toluene was also checked; for example, the I100Q mutant formed 22% o-cresol, 44% m-cresol, and 34% p-cresol, whereas wild-type ToMO formed 32% o-cresol, 21% m-cresol, and 47% p-cresol.  相似文献   

15.
Toluene-o-xylene monooxygenase (ToMO) from Pseudomonas stutzeri OX1 oxidizes toluene to 3- and 4-methylcatechol and oxidizes benzene to form phenol; in this study ToMO was found to also form catechol and 1,2,3-trihydroxybenzene (1,2,3-THB) from phenol. To synthesize novel dihydroxy and trihydroxy derivatives of benzene and toluene, DNA shuffling of the alpha-hydroxylase fragment of ToMO (TouA) and saturation mutagenesis of the TouA active site residues I100, Q141, T201, and F205 were used to generate random mutants. The mutants were initially identified by screening with a rapid agar plate assay and then were examined further by high-performance liquid chromatography and gas chromatography. Several regiospecific mutants with high rates of activity were identified; for example, Escherichia coli TG1/pBS(Kan)ToMO expressing the F205G TouA saturation mutagenesis variant formed 4-methylresorcinol (0.78 nmol/min/mg of protein), 3-methylcatechol (0.25 nmol/min/mg of protein), and methylhydroquinone (0.088 nmol/min/mg of protein) from o-cresol, whereas wild-type ToMO formed only 3-methylcatechol (1.1 nmol/min/mg of protein). From o-cresol, the I100Q saturation mutagenesis mutant and the M180T/E284G DNA shuffling mutant formed methylhydroquinone (0.50 and 0.19 nmol/min/mg of protein, respectively) and 3-methylcatechol (0.49 and 1.5 nmol/min/mg of protein, respectively). The F205G mutant formed catechol (0.52 nmol/min/mg of protein), resorcinol (0.090 nmol/min/mg of protein), and hydroquinone (0.070 nmol/min/mg of protein) from phenol, whereas wild-type ToMO formed only catechol (1.5 nmol/min/mg of protein). Both the I100Q mutant and the M180T/E284G mutant formed hydroquinone (1.2 and 0.040 nmol/min/mg of protein, respectively) and catechol (0.28 and 2.0 nmol/min/mg of protein, respectively) from phenol. Dihydroxybenzenes were further oxidized to trihydroxybenzenes with different regiospecificities; for example, the I100Q mutant formed 1,2,4-THB from catechol, whereas wild-type ToMO formed 1,2,3-THB (pyrogallol). Regiospecific oxidation of the natural substrate toluene was also checked; for example, the I100Q mutant formed 22% o-cresol, 44% m-cresol, and 34% p-cresol, whereas wild-type ToMO formed 32% o-cresol, 21% m-cresol, and 47% p-cresol.  相似文献   

16.
Ullrich R  Hofrichter M 《FEBS letters》2005,579(27):6247-6250
The mushroom Agrocybe aegerita secretes a peroxidase (AaP) that catalyzes halogenations and hydroxylations. Phenol was brominated to 2- and 4-bromophenol (ratio 1:4) and chlorinated to a lesser extent to 2-chlorophenol. The purified enzyme was found to oxidize toluene via benzyl alcohol and benzaldehyde into benzoic acid. A second fraction of toluene was hydroxylated to give p-cresol as well as o-cresol and methyl-p-benzoquinone. The UV-Vis absorption spectrum of purified AaP showed high similarity to a resting state cytochrome P450 with the Soret band at 420 nm and additional maxima at 278, 358, 541 and 571 nm; the AaP CO-complex had a distinct absorption maximum at 445 nm that is characteristic for heme-thiolate proteins. AaP regioselectively hydroxylated naphthalene to 1-naphthol and traces of 2-naphthol (ratio 36:1). H2O2 was necessarily required for AaP function and hence the hydroxylations catalyzed by AaP can be designated as peroxygenation and the enzyme as an extracellular peroxygenase.  相似文献   

17.
Of the four investigated Rhodococcus strains (R. rhodochrous172, R. opacus 4a and 557, and R. rhodnii 135), the first three strains were found to be able to completely transform fluorene when it was present in the medium as the sole source of carbon at a concentration of 12–25 mg/l. At a fluorene concentration of 50–100 mg/l in the medium, the rhodococci transformed 50% of the substrate in 14 days. The addition of casamino acids and sucrose (1–5 g/l) stimulated fluorene transformation, so that R. rhodochrous 172 could completely transform it in 2–5 days. Nine intermediates of fluorene transformation were isolated, purified, and structurally characterized. It was found that R. rhodnii 135 and R. opacus strains 4a and 557 hydroxylated fluorene with the formation of 2-hydroxyfluorene and 2,7-dihydroxyfluorene. R. rhodochrous 172 transformed fluorene via two independent pathways to a greater degree than the other rhodococci studied.  相似文献   

18.
Mitchell KH  Studts JM  Fox BG 《Biochemistry》2002,41(9):3176-3188
Toluene 4-monooxygenase (T4MO) is a diiron hydroxylase that exhibits high regiospecificity for para hydroxylation. This fidelity provides the basis for an assessment of the interplay between active site residues and protein complex formation in producing an essential biological outcome. The function of the T4MO catalytic complex (hydroxylase, T4moH, and effector protein T4moD) is evaluated with respect to effector protein concentration, the presence of T4MO electron-transfer components (Rieske ferredoxin, T4moC, and NADH oxidoreductase), and use of mutated T4moH isoforms with different hydroxylation regiospecificities. Steady-state kinetic analyses indicate that T4moC and T4moD form complexes of similar affinity with T4moH. At low T4moD concentrations, the steady-state hydroxylation rate is linearly dependent on T4moD-T4moH complex formation, whereas regiospecificity and the coupling efficiency between NADH consumption and hydroxylation are associated with intrinsic properties of the T4moD-T4moH complex. The optimized complex gives both efficient coupling and high regiospecificity with p-cresol representing >96% of total products from toluene. Similar coupling and regiospecificity for para hydroxylation are obtained with T3buV (an effector protein from a toluene 3-monooxygenase), demonstrating that effector protein binding does not uniquely determine or alter the regiospecificity of toluene hydroxylation. The omission of T4moD causes an approximately 20-fold decrease in hydroxylation rate, nearly complete uncoupling, and a decrease in regiospecificity so that p-cresol represents approximately 60% of total products. Similar shifts in regiospecificity are observed in oxidations of alternative substrates in the absence or upon the partial removal of either T4moD or T3buV from toluene oxidations. The mutated T4moH isoforms studied have apparent V(max)/K(M) specificities differing by approximately 2-4-fold and coupling efficiencies ranging from 88% to 95%, indicating comparable catalytic function, but also exhibit unique regiospecificity patterns for all substrates tested, suggesting unique substrate binding preferences within the active site. The G103L isoform has enhanced selectivity for ortho hydroxylation with all substrates tested except nitrobenzene, which gives only m-nitrophenol. The regiospecificity of the G103L isoform is comparable to that observed from naturally occurring variants of the toluene/benzene/o-xylene monooxygenase subfamily. Evolutionary and mechanistic implications of these findings are considered.  相似文献   

19.
Abstract Three bacterial isolates which appeared to use the insecticide, carbaryl (1-naphthyl, N -methyl-carbamate) as their sole carbon and nitrogen sources were originally selected from garden soil. Only one isolate, Pseudomonas sp. (NCIB 12043) could metabolise carbaryl rapidly to 1-naphthol and methylamine. The other two isolates, Pseudomonas sp. (NCIB 12042) and Rhodococcus sp. (NCIB 12038) relied on slow chemical hydrolysis of carbaryl to 1-naphthol and methylamine. All three isolates used 1-naphthol as their sole carbon source; however, their ability to use naphthalene and a range of mono- and dihydroxy-substituted naphthalene compounds varied. NCIB 12038 and NCIB 12043 showed little or no growth on naphthalene, 2,3-dihydroxynaphthalene or 1,3-dihydroxynaphthalene as sole carbon sources and their 1-naphthol oxygenases had little activity with these substrates. In contrast, NCIB 12042 could use these compounds as sole carbon sources and its 1-naphthol oxygenase also showed activity with them. We conclude that 1-naphthol oxygenase from NCIB 12042 is a relatively non-specific dioxygenase, whereas the 1-naphthol oxygenases from NCIB 12038 and NCIB 12043 are relatively specific monooxygenases requiring hydroxylated naphthalene compounds as substrates.  相似文献   

20.
Decomposition of β-Naphthol by a Soil Pseudomonad   总被引:1,自引:1,他引:0  
Summary: A pseudomonad resembling Pseudomonas fluorescens , which grows with β-naphthol as sole source of carbon, was isolated from soil. It did not grow on either naphthalene, α-naphthol, 1,2- or 2,3 dihydroxynaphthalene. Phenol, benzoic acid, o-, p - and (to a small extent) m -hydroxybenzoic acids supported growth of the organism. A maroon coloured substance was produced from β-naphthol in cultures and by washed organisms. β-Naphthol oxidation depended on an induced enzyme system. β-Naphthol-grown organisms oxidized β-naphthol and 2,3- and 2,6-dihydroxynaphthalene immediately and several mono- and di-hydroxybenzoic acids, including salicylic acid, only after a lag. 2,3-Dihydroxynaphthalene may be a metabolite of β-naphthol.  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