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 455 毫秒
1.
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   

2.
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   

3.
After discovering that toluene 4-monooxygenase (T4MO) of Pseudomonas mendocina KR1 oxidizes nitrobenzene to 4-nitrocatechol, albeit at a very low rate, this reaction was improved using directed evolution and saturation mutagenesis. Screening 550 colonies from a random mutagenesis library generated by error-prone PCR of tmoAB using Escherichia coli TG1/pBS(Kan)T4MO on agar plates containing nitrobenzene led to the discovery of nitrocatechol-producing mutants. One mutant, NB1, contained six amino acid substitutions (TmoA Y22N, I84Y, S95T, I100S, S400C; TmoB D79N). It was believed that position I100 of the alpha subunit of the hydroxylase (TmoA) is the most significant for the change in substrate reactivity due to previous results in our lab with a similar enzyme, toluene ortho-monooxygenase of Burkholderia cepacia G4. Saturation mutagenesis at this position resulted in the generation of two more nitrocatechol mutants, I100A and I100S; the rate of 4-nitrocatechol formation by I100A was more than 16 times higher than that of wild-type T4MO at 200 microM nitrobenzene (0.13 +/- 0.01 vs. 0.008 +/- 0.001 nmol/min.mg protein). HPLC and mass spectrometry analysis revealed that variants NB1, I100A, and I100S produce 4-nitrocatechol via m-nitrophenol, while the wild-type produces primarily p-nitrophenol and negligible amounts of nitrocatechol. Relative to wild-type T4MO, whole cells expressing variant I100A convert nitrobenzene into m-nitrophenol with a Vmax of 0.61 +/- 0.037 vs. 0.16 +/- 0.071 nmol/min.mg protein and convert m-nitrophenol into nitrocatechol with a Vmax of 3.93 +/- 0.26 vs. 0.58 +/- 0.033 nmol/min.mg protein. Hence, the regiospecificity of nitrobenzene oxidation was changed by the random mutagenesis, and this led to a significant increase in 4-nitrocatechol production. The regiospecificity of toluene oxidation was also altered, and all of the mutants produced 20% m-cresol and 80% p-cresol, while the wild-type produces 96% p-cresol. Interestingly, the rate of toluene oxidation (the natural substrate of the enzyme) by I100A was also higher by 65% (7.2 +/- 1.2 vs. 4.4 +/- 0.3 nmol/min mg protein). Homology-based modeling of TmoA suggests reducing the size of the side chain of I100 leads to an increase in the width of the active site channel, which facilitates access of substrates and promotes more flexible orientations.  相似文献   

4.
The regiospecific oxidation of the polycyclic aromatic hydrocarbons naphthalene and fluorene was examined with Escherichia coli strains expressing wildtype toluene 4-monooxygenase (T4MO) from Pseudomonas mendocina KR1, toluene para-monooxygenase (TpMO) from Ralstonia pickettii PKO1, toluene ortho-monooxygenase (TOM) from Burkholderia cepacia G4, and toluene/ortho-xylene monooxygenase (ToMO) from P. stutzeri OX1. T4MO oxidized toluene (12.1+/-0.8 nmol/min/mg protein at 109 microM), naphthalene (7.7+/-1.5 nmol/min/mg protein at 5 mM), and fluorene (0.68+/-0.04 nmol/min/mg protein at 0.2 mM) faster than the other wildtype enzymes (2-22-fold) and produced a mixture of 1-naphthol (52%) and 2-naphthol (48%) from naphthalene, which was successively transformed to a mixture of 2,3-, 2,7-, 1,7-, and 2,6-dihydroxynaphthalenes (7%, 10%, 20%, and 63%, respectively). TOM and ToMO made 1,7-dihydroxynaphthalene from 1-naphthol, and ToMO made a mixture of 2,3-, 2,6-, 2,7-, and 1,7-dihydroxynaphthalene (26%, 22%, 1%, and 44%, respectively) from 2-naphthol. TOM had no activity on 2-naphthol, and T4MO had no activity on 1-naphthol. To take advantage of the high activity of wildtype T4MO but to increase its regiospecificity on naphthalene, seven engineered enzymes containing mutations in T4MO alpha hydroxylase TmoA were examined; the selectivity for 2-naphthol by T4MO I100A, I100S, and I100G was enhanced to 88-95%, and the selectivity for 1-naphthol was enhanced to 87% and 99% by T4MO I100L and G103S/A107G, respectively, while high oxidation rates were maintained except for G103S/A107G. Therefore, the regiospecificity for naphthalene oxidation was altered to practically pure 1-naphthol or 2-naphthol. All four wildtype monooxygenases were able to oxidize fluorene to different monohydroxylated products; T4MO oxidized fluorene successively to 3-hydroxyfluorene and 3,6-dihydroxyfluorene, which was confirmed by gas chromatography-mass spectrometry and 1H nuclear magnetic resonance analysis. TOM and its variant TomA3 V106A oxidize fluorene to a mixture of 1-, 2-, 3-, and 4-hydroxyfluorene. This is the first report of using enzymes to synthesize 1-, 3-, and 4-hydroxyfluorene, and 3,6-dihydroxyfluorene from fluorene as well as 2-naphthol and 2,6-dihydroxynaphthalene from naphthalene.  相似文献   

5.
Saturation mutagenesis of the 2,4-dinitrotoluene dioxygenase (DDO) of Burkholderia cepacia R34 at position valine 350 of the DntAc alpha-subunit generated mutant V350F with significantly increased activity towards o-nitrophenol (47 times), m-nitrophenol (34 times), and o-methoxyphenol (174 times) as well as an expanded substrate range that now includes m-methoxyphenol, o-cresol, and m-cresol (wild-type DDO had no detectable activity for these substrates). Another mutant, V350M, also displays increased activity towards o-nitrophenol (20 times) and o-methoxyphenol (162 times) as well as novel activity towards o-cresol. Products were synthesized using whole Escherichia coli TG1 cells expressing the recombinant R34 dntA loci from pBS(Kan)R34, and the initial rates of product formation were determined at 1 mM substrate by reverse-phase high-pressure liquid chromatography. V350F produced both nitrohydroquinone at a rate of 0.75 +/- 0.15 nmol/min/mg of protein and 3-nitrocatechol at a rate of 0.069 +/- 0.001 nmol/min/mg of protein from o-nitrophenol, 4-nitrocatechol from m-nitrophenol at 0.29 +/- 0.02 nmol/min/mg of protein, methoxyhydroquinone from o-methoxyphenol at 2.5 +/- 0.6 nmol/min/mg of protein, methoxyhydroquinone from m-methoxyphenol at 0.55 +/- 0.02 nmol/min/mg of protein, both methylhydroquinone at 1.52 +/- 0.02 nmol/min/mg of protein and 2-hydroxybenzyl alcohol at 0.74 +/- 0.05 nmol/min/mg of protein from o-cresol, and methylhydroquinone at 0.43 +/- 0.1 nmol/min/mg of protein from m-cresol. V350M produced both nitrohydroquinone at a rate of 0.33 nmol/min/mg of protein and 3-nitrocatechol at 0.089 nmol/min/mg of protein from o-nitrophenol, methoxyhydroquinone from o-methoxyphenol at 2.4 nmol/min/mg of protein, methylhydroquinone at 1.97 nmol/min/mg of protein and 2-hydroxybenzyl alcohol at 0.11 nmol/min/mg of protein from o-cresol. The DDO variants V350F and V350M also exhibited 10-fold-enhanced activity towards naphthalene (8 +/- 2.6 nmol/min/mg of protein), forming (1R,2S)-cis-1,2-dihydro-1,2-dihydroxynaphthalene. Hence, mutagenesis of wild-type DDO through active-site engineering generated variants with relatively high rates toward a previously uncharacterized class of substituted phenols for the nitroarene dioxygenases; seven previously uncharacterized substrates were evaluated for wild-type DDO, and four novel monooxygenase-like products were found for the DDO variants V350F and V350M (methoxyhydroquinone, methylhydroquinone, 2-hydroxybenzyl alcohol, and 3-nitrocatechol).  相似文献   

6.
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   

7.
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   

8.
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   

9.
Directed evolution of toluene ortho-monooxygenase (TOM) of Burkholderia cepacia G4 previously created the hydroxylase α-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 (Vmax of 9.3 versus 4.5 nmol·min−1·mg of protein−1 and unchanged Km), 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 μM 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 (Vmax of 2.61 versus 0.95 nmol·min−1·mg of protein−1 and unchanged Km) 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   

10.
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   

11.
Pseudomonas cepacia G4 possesses a novel pathway of toluene catabolism that is shown to be responsible for the degradation of trichloroethylene (TCE). This pathway involves conversion of toluene via o-cresol to 3-methylcatechol. In order to determine the enzyme of toluene degradation that is responsible for TCE degradation, chemically induced mutants, blocked in the toluene ortho-monooxygenase (TOM) pathway of G4, were examined. Mutants of the phenotypic class designated TOM A- were all defective in their ability to oxidize toluene, o-cresol, m-cresol, and phenol, suggesting that a single enzyme is responsible for conversion of these compounds to their hydroxylated products (3-methylcatechol from toluene, o-cresol, and m-cresol and catechol from phenol) in the wild type. Mutants of this class did not degrade TCE. Two other mutant classes which were blocked in toluene catabolism, TOM B-, which lacked catechol-2,3-dioxygenase, and TOM C-, which lacked 2-hydroxy-6-oxoheptadienoic acid hydrolase activity, were fully capable of TCE degradation. Therefore, TCE degradation is directly associated with the monooxygenation capability responsible for toluene, cresol, and phenol hydroxylation.  相似文献   

12.
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   

13.
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   

14.
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   

15.
A mixed culture of nitrate-reducing bacteria degraded o-cresol in the presence of toulene as a primary growth substrate. No degradation of o-cresol was observed in the absence of toluene or when the culture grew on p-cresol and 2,4-dimethylphenol. In batch cultures, the degradation of o-cresol started after toluene was degraded to below 0.5 to 1.0 mg/liter but continued only for about 3 to 5 days after the depletion of toluene since the culture had a limited capacity for o-cresol degradation once toluene was depleted. The total amount of o-cresol degraded was proportional to the amount of toluene metabolized, with an average yield of 0.47 mg of o-cresol degraded per mg of toluene metabolized. Experiments with [ring-U-14C]o-cresol indicated that about 73% of the carbon from degraded o-cresol was mineralized to CO2 and about 23% was assimilated into biomass after the transient accumulation of unidentified water-soluble intermediates. A mathematical model based on a simplified Monod equation is used to describe the kinetics of o-cresol degradation. In this model, the biomass activity toward o-cresol is assumed to decay according to first-order kinetics once toluene is depleted. On the basis of nonlinear regression of the data, the maximum specific rate of o-cresol degradation was estimated to be 0.4 mg of o-cresol per mg of biomass protein per h, and the first-order decay constant for o-cresol-degrading biomass activity was estimated to be 0.15 h-1.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
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   

17.
Aerobic, co-metabolic bioremediation of trichloroethylene (TCE), cis-1,2-dichloroethylene (cis-DCE) and other chlorinated ethenes with monooxygenase-expressing microorganisms is limited by the toxic epoxides produced as intermediates. A recombinant Escherichia coli strain less sensitive to the toxic effects of cis-DCE, TCE and trans-1,2-dichloroethylene (trans-DCE) degradation has been created by engineering a novel pathway consisting of eight genes including a DNA-shuffled toluene ortho-monooxygenase from Burkholderia cepacia G4 (TOM-Green), a newly discovered glutathione S-transferase (GST) from RhodococcusAD45 (IsoILR1), found to have activity towards epoxypropane and cis-DCE epoxide, and an overexpressed E. coli mutant gamma-glutamylcysteine synthetase (GSHI*). Along with IsoILR1, another new RhodococcusAD45 GST, IsoILR2, was cloned that lacks activity towards cis-DCE epoxide and differs from IsoILR1 by nine amino acids. The recombinant strain in which TOM-Green and IsoILR1 were co-expressed on separate plasmids degraded 1.9-fold more cis-DCE compared with a strain that lacked IsoILR1. In the presence of IsoILR1 and TOM-Green, the addition of GSH1* resulted in a sevenfold increase in the intracellular GSH concentration and a 3.5-fold improvement in the cis-DCE degradation rate based on chloride released (2.1 +/- 0.1 versus 0.6 +/- 0.1 nmol min(-1) mg(-1) protein at 540 microM), a 1.8-fold improvement in the trans-DCE degradation rate (1.29 +/- 0.03 versus 0.71 +/- 0.04 nmol x min(-1) mg(-1) protein at 345 microM) and a 1.7-fold improvement in the TCE degradation rate (6.8 +/- 0.24 versus 4.1 +/- 0.16 nmol x min(-1) mg(-1) protein at 339 microM). For cis-DCE degradation with TOM-Green (based on substrate depletion), V(max) was 27 nmol x min(-1) mg(-1) protein with both IsoILR1 and GSHI* expressed compared with V(max) = 10 nmol x min(-1) mg(-1) protein for the GST(-)GSHI*(-) strain. In addition, cells expressing IsoILR1 and GSHI* grew 78% faster in rich medium than a strain lacking these two heterologous genes.  相似文献   

18.
Enantiopure sulfoxides are valuable asymmetric starting materials and are important chiral auxiliaries in organic synthesis. Toluene monooxygenases (TMOs) have been shown previously to catalyze regioselective hydroxylation of substituted benzenes and phenols. Here we show that TMOs are also capable of performing enantioselective oxidation reactions of aromatic sulfides. Mutagenesis of position V106 in the α-hydroxylase subunit of toluene ortho-monooxygenase (TOM) of Burkholderia cepacia G4 and the analogous position I100 in toluene 4-monooxygenase (T4MO) of Pseudomonas mendocina KR1 improved both rate and enantioselectivity. Variant TomA3 V106M of TOM oxidized methyl phenyl sulfide to the corresponding sulfoxide at a rate of 3.0 nmol/min/mg protein compared with 1.6 for the wild-type enzyme, and the enantiomeric excess (pro-S) increased from 51% for the wild type to 88% for this mutant. Similarly, T4MO variant TmoA I100G increased the wild-type oxidation rate by 1.7-fold, and the enantiomeric excess rose from 86% to 98% (pro-S). Both wild-type enzymes showed lower activity with methyl para-tolyl sulfide as a substrate, but the improvement in the activity and enantioselectivity of the mutants was more dramatic. For example, T4MO variant TmoA I100G oxidized methyl para-tolyl sulfide 11 times faster than the wild type did and changed the selectivity from 41% pro-R to 77% pro-S. A correlation between regioselectivity and enantioselectivity was shown for TMOs studied in this work. Using in silico homology modeling, it is shown that residue I100 in T4MO aids in steering the substrate into the active site at the end of the long entrance channel. It is further hypothesized that the main function of V106 in TOM is the proper positioning or docking of the substrate with respect to the diiron atoms. The results from this work suggest that when the substrate is not aligned correctly in the active site, the oxidation rate is decreased and enantioselectivity is impaired, resulting in products with both chiral configurations.  相似文献   

19.
4-Methyl-5-nitrocatechol (4M5NC) monooxygenase (DntB) from Burkholderia sp. strain DNT catalyzes the second step of 2,4-dinitrotoluene degradation by converting 4M5NC to 2-hydroxy-5-methylquinone with the concomitant removal of the nitro group. DntB is a flavoprotein that has a very narrow substrate range. Here, error-prone PCR was used to create variant DntB M22L/L380I, which accepts the two new substrates 4-nitrophenol (4NP) and 3-methyl-4-nitrophenol (3M4NP). At 300 microM of 4NP, the initial rate of the variant expressing M22L/L380I enzyme (39 +/- 6 nmol/min/mg protein) was 10-fold higher than that of the wild-type enzyme (4 +/- 2 nmol/min/mg protein). The values of kcat/Km of the purified wild-type DntB enzyme and purified variant M22L/L380I were 40 and 450 (s(-1) M(-1)), respectively, which corroborates that the variant M22L/L380I enzyme has 11-fold-higher efficiency than the wild-type enzyme for 4NP degradation. In addition, the variant M22L/L380I enzyme has fourfold-higher activity toward 3M4NP; at 300 microM, the initial nitrite release rate of M22L/L380I enzyme was 17 +/- 4 nmol/min/mg protein, while that of the wild-type enzyme was 4.4 +/- 0.7 nmol/min/mg protein. Saturation mutagenesis was also used to further investigate the role of the individual amino acid residues at positions M22, L380, and M22/L380 simultaneously. Mutagenesis at the individual positions M22L and L380I did not show appreciable enhancement in 4NP activity, which suggested that these two sites should be mutated together; simultaneous saturation mutagenesis led to the identification of the variant M22S/L380V, with 20% enhanced degradation of 4NP compared to the variant M22L/L380I. This is the first report of protein engineering for nitrite removal by a flavoprotein.  相似文献   

20.
DNA shuffling and saturation mutagenesis of positions F108, L190, I219, D235, and C248 were used to generate variants of the epoxide hydrolase of Agrobacterium radiobacter AD1 (EchA) with enhanced enantioselectivity and activity for styrene oxide and enhanced activity for 1,2-epoxyhexane and epoxypropane. EchA variant I219F has more than fivefold-enhanced enantioselectivity toward racemic styrene oxide, with the enantiomeric ratio value (E value) for the production of (R)-1-phenylethane-1,2-diol increased from 17 for the wild-type enzyme to 91, as well as twofold-improved activity for the production of (R)-1-phenylethane-1,2-diol (1.96 +/- 0.09 versus 1.04 +/- 0.07 micromol/min/mg for wild-type EchA). Computer modeling indicated that this mutation significantly alters (R)-styrene oxide binding in the active site. Another three variants from EchA active-site engineering, F108L/C248I, I219L/C248I, and F108L/I219L/C248I, also exhibited improved enantioselectivity toward racemic styrene oxide in favor of production of the corresponding diol in the (R) configuration (twofold enhancement in their E values). Variant F108L/I219L/C248I also demonstrated 10-fold- and 2-fold-increased activity on 5 mM epoxypropane (24 +/- 2 versus 2.4 +/- 0.3 micromol/min/mg for the wild-type enzyme) and 5 mM 1,2-epoxyhexane (5.2 +/- 0.5 versus 2.6 +/- 0.0 micromol/min/mg for the wild-type enzyme). Both variants L190F (isolated from a DNA shuffling library) and L190Y (created from subsequent saturation mutagenesis) showed significantly enhanced activity for racemic styrene oxide hydrolysis, with 4.8-fold (8.6 +/- 0.3 versus 1.8 +/- 0.2 micromol/min/mg for the wild-type enzyme) and 2.7-fold (4.8 +/- 0.8 versus 1.8 +/- 0.2 micromol/min/mg for the wild-type enzyme) improvements, respectively. L190Y also hydrolyzed 1,2-epoxyhexane 2.5 times faster than the wild-type enzyme.  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