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 15 毫秒
1.
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   

2.
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   

3.
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   

4.
2,4-Dinitrotoluene (2,4-DNT) and 2,6-DNT are priority pollutants, and 2,4-DNT dioxygenase of Burkholderia sp. strain DNT (DDO) catalyzes the initial oxidation of 2,4-DNT to form 4-methyl-5-nitrocatechol and nitrite but has significantly less activity on other dinitrotoluenes and nitrotoluenes (NT). Hence, oxidation of 2,3-DNT, 2,4-DNT, 2,5-DNT, 2,6-DNT, 2NT, and 4NT were enhanced here by performing saturation mutagenesis on codon I204 of the alpha subunit (DntAc) of DDO and by using a membrane agar plate assay to detect catechol formation. Rates of degradation were quantified both by the formation of nitrite and by the formation of the intermediates with high performance liquid chromatography. The degradation of both 2,3-DNT and 2,5-DNT were achieved for the first time (no detectable activity with the wild-type enzyme) using whole Escherichia coli TG1 cells expressing DDO variants DntAc I204L and I204Y (0.70 +/- 0.03 and 0.22 +/- 0.02 nmol/min/mg protein for 2,5-DNT transformation, respectively). DDO DntAc variant I204L also transformed both 2,6-DNT and 2,4-DNT 2-fold faster than wild-type DDO (0.8 +/- 0.6 nmol/min/mg protein and 4.7 +/- 0.5 nmol/min/mg protein, respectively). Moreover, the activities of DDO for 2NT and 4NT were also enhanced 3.5-fold and 8-fold, respectively. Further, DntAc variant I204Y was also discovered with comparable rate enhancements for the substrates 2,4-DNT, 2,6-DNT, and 2NT but not 4NT. Sequencing information obtained during this study indicated that the 2,4-DNT dioxygenases of Burkholderia sp. strain DNT and B. cepacia R34 are more closely related than originally reported. This is the first report of engineering an enzyme for enhanced degradation of nitroaromatic compounds and the first report of degrading 2,5-DNT.  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.
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   

7.
Plant constituents such as terpenes are major constituents of the essential oil in Eucalyptus sp. 1,8-Cineole and p-cymene (Terpenes present in high amounts in Eucalyptus leaves) are potential substrates for the CYP family of enzymes. We have investigated tolbutamide hydroxylase as a probe substrate reaction in both koala and terpene pretreated and control brushtail possum liver microsomes and examined inhibition of this reaction by Eucalyptus terpenes. The specific activity determined for tolbutamide hydroxylase in the terpene treated brushtails was significantly higher than that for the control animals (1865+/-334 nmol/mg microsomal protein per min versus 895+/-27 nmol/mg microsomal protein per min). The activity determined in koala microsomes was 8159+/-370 nmol/mg microsomal protein per min. Vmax values and Km values for the terpene treated possum, control, possum and koala were 1932-2225 nmol/mg microsomal protein per min and 0.80 0.81 mM; 1406-1484 nmol/mg microsomal protein per min and 0.87-0.92 mM and 5895-6403 nmol/mg microsomal protein per min and 0.067-0.071 mM, respectively. Terpenes were examined as potential inhibitors of tolbutamide hydroxylase activity. 1,8-Cineole was found to be a competitive inhibitor for the enzyme responsible for tolbutamide hydroxylation (Ki 15 microM) in the possum. In koala liver microsomes stimulation of tolbutamide hydroxylase activity was observed when concentrations of cineole were increased. Therefore, although inhibition was observed, the type of inhibition could not be determined.  相似文献   

8.
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   

9.
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   

10.
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   

11.
Metabolically engineered Escherichia coli has previously been used to degrade cis-1,2-dichloroethylene (cis-DCE). The strains express the six genes of an evolved toluene ortho-monooxygenase from Burkholderia cepacia G4 (TOM-Green, which formed a reactive epoxide) with either (1) gamma-glutamylcysteine synthetase (GSHI, which forms glutathione) and the glutathione S-transferase IsoILR1 from Rhodococcus AD45 (which adds glutathione to the reactive cis-DCE epoxide) or (2) with an evolved epoxide hydrolase from Agrobacterium radiobacter AD1 (EchA F108L/I219L/C248I which converts the reactive cis-DCE epoxide to a diol). Here, the impact of this metabolic engineering for bioremediation was assessed by investigating the changes in the proteome through a quantitative shotgun proteomics technique (iTRAQ) by tracking the changes due to the sequential addition of TOM-Green, IsoILR1, and GSHI and due to adding the evolved EchA versus the wild-type enzyme to TOM-Green. For the TOM-Green/EchA system, 8 proteins out of 268 identified proteins were differentially expressed in the strain expressing EchA F108L/I219L/C248I relative to wild-type EchA (e.g., EchA, protein chain elongation factor EF-Ts, 50S ribosomal subunits L7/L12/L32/L29, cysteine synthase A, glycerophosphodiester phosphodiesterase, iron superoxide dismutase). For the TOM-Green/IsoILR1/GSHI system, the expression level of 49 proteins was changed out of 364 identified proteins. The induced proteins due to the addition of TOM-Green, IsoILR1, and GSHI were involved in the oxidative defense mechanism, pyruvate metabolism, and glutathione synthesis (e.g., 30S ribosomal subunit proteins S3 and S16, 50S ribosomal subunit protein L20, alkyl hydroperoxide reductase, lactate dehydrogenase, acetate kinase, cysteine synthase A). Enzymes involved in indole synthesis, fatty acid synthesis, gluconeogenesis, and the tricarboxylic acid cycle were repressed (e.g., tryptophanase, acetyl-CoA carboxylase, phosphoenolpyruvate carboxykinase, malate dehydrogenase). Hence, the metabolic engineering that leads to enhanced aerobic degradation of 1 mM cis-DCE (2.4-4-fold more chloride ions released) and reduced toxicity from cis-DCE epoxide results in enhanced synthesis of glutathione coupled with an induced stress response as well as repression of fatty acid synthesis, gluconeogenesis, and the tricarboxylic acid cycle.  相似文献   

12.
Intact cells of Pseudomonas cepacia G4 completely degraded trichloroethylene (TCE) following growth with phenol. Degradation kinetics were determined for both phenol, used to induce requisite enzymes, and TCE, the target substrate. Apparent Ks and Vmax values for degradation of phenol by cells were 8.5 microM and 466 nmol/min per mg of protein, respectively. At phenol concentrations greater than 50 microM, phenol degradation was inhibited, yielding an apparent second-order inhibitory value, KSI, of 0.45 mM as modeled by the Haldane expression. A partition coefficient for TCE was determined to be 0.40 +/- 0.02, [TCEair]/[TCEwater], consistent with Henry's law. To eliminate experimental problems associated with TCE volatility and partitioning, a no-headspace bottle assay was developed, allowing for direct and accurate determinations of aqueous TCE concentration. By this assay procedure, apparent Ks and Vmax values determined for TCE degradation by intact cells were 3 microM and 8 nmol/min per mg of protein, respectively. Following a transient lag period, P. cepacia G4 degraded TCE at concentrations of at least 300 microM with no apparent retardation in rate. Consistent with Ks values determined for degradation, TCE significantly inhibited phenol degradation.  相似文献   

13.
Intact cells of Pseudomonas cepacia G4 completely degraded trichloroethylene (TCE) following growth with phenol. Degradation kinetics were determined for both phenol, used to induce requisite enzymes, and TCE, the target substrate. Apparent Ks and Vmax values for degradation of phenol by cells were 8.5 microM and 466 nmol/min per mg of protein, respectively. At phenol concentrations greater than 50 microM, phenol degradation was inhibited, yielding an apparent second-order inhibitory value, KSI, of 0.45 mM as modeled by the Haldane expression. A partition coefficient for TCE was determined to be 0.40 +/- 0.02, [TCEair]/[TCEwater], consistent with Henry's law. To eliminate experimental problems associated with TCE volatility and partitioning, a no-headspace bottle assay was developed, allowing for direct and accurate determinations of aqueous TCE concentration. By this assay procedure, apparent Ks and Vmax values determined for TCE degradation by intact cells were 3 microM and 8 nmol/min per mg of protein, respectively. Following a transient lag period, P. cepacia G4 degraded TCE at concentrations of at least 300 microM with no apparent retardation in rate. Consistent with Ks values determined for degradation, TCE significantly inhibited phenol degradation.  相似文献   

14.
The activity of phosphoglucose isomerase, its kinetic properties, and the effect of 6-phosphogluconate on its activity in the forward (glucose 6-phosphate----fructose 6-phosphate) and the reverse (fructose 6-phosphate----glucose 6-phosphate) reactions were determined in adult rat brain in vitro. The activity of phosphoglucose isomerase (in nmol/min/mg of whole brain protein) was 1,865 +/- 20 in the forward reaction and 1,756 +/- 32 in the reverse reaction at pH 7.5. It was 1,992 +/- 28 and 2,620 +/- 46, respectively, at pH 8.5. The apparent Km and Vmax of phosphoglucose isomerase were 0.593 +/- 0.031 mM and 2,291 +/- 61 nmol/min/mg of protein, respectively, for glucose 6-phosphate and 0.095 +/- 0.013 mM and 2,035 +/- 98 nmol/min/mg of protein, respectively, for fructose 6-phosphate. The activity of phosphoglucose isomerase was inhibited intensely and competitively by 6-phosphogluconate, with an apparent Ki of 0.048 +/- 0.005 mM for glucose 6-phosphate and 0.042 +/- 0.004 mM for fructose 6-phosphate as the substrate. With glucose 6-phosphate as the substrate, at concentrations from 0.05 to 0.5 mM, the activity of the enzyme was inhibited completely in the presence of 0.5-2.0 mM 6-phosphogluconate. With 0.05-0.2 mM fructose 6-phosphate as the substrate, it was inhibited greater than or equal to 85% at the same concentrations of the inhibitor. No significant changes were observed in the values of Km, Vmax, and Ki for phosphoglucose isomerase in the brain of 6-aminonicotinamide-treated rats.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
We examined the in vitro effects of tert-butylhydroperoxide (tBu-OOH) in human bronchial muscle. tert-Butylhydroperoxide produced concentration-dependent contractions of bronchial rings (maximum effect was 56.5 +/- 9.6% of contraction by 1 mM acetylcholine; effective concentration 50% was approximately 100 microM). tert-Butylhydroperoxide (0.5 mM)-induced contraction was enhanced by epithelial removal but abolished by indomethacin (cyclooxygenase inhibitor) and zileuton (lipoxygenase inhibitor). tert-Butylhydroperoxide produced a transient rise in intracellular calcium in human cultured airway smooth muscle cells (HCASMC). The bronchial reactivity to acetylcholine and histamine was not altered by tBu-OOH. In HCASMC, tBu-OOH (0.5 mM, 30 min) increased malondialdehyde levels (MDA; from 7.80 +/- 0.83 to 26.82 +/- 1.49 nmol mg(-1) protein), accompanied by a decrease of reduced glutathione (GSH; from 16.7 +/- 2.6 to 6.9 +/- 1.9 nmol mg(-1) protein) and an increase of oxidized glutathione (from 0.09 +/- 0.03 to 0.18 +/- 0.03 nmol mg(-1) protein). N-acetylcysteine (0.3 mM) inhibited by approximately 60% the bronchial contraction resulting from tBu-OOH (0.5 mM) and protected cultured cells exposed to tBu-OOH (MDA was lowered to 19.51 +/- 1.19 nmol mg(-1) protein, and GSH content was replenished). In summary, tBu-OOH caused contraction of human bronchial muscle mediated by release of cyclo-oxygenase and lipoxygenase products without producing airways hyperreactivity. N-acetylcysteine decreases tBu-OOH-induced contraction and protects human cultured airway smooth muscle cells exposed to tBu-OOH.  相似文献   

16.
Trichloroethylene (TCE) was reductively dechlorinated to cis-dichloroethylene, trans-dichloroethylene, 1,1-dichloroethylene, vinyl chloride, and ethylene by the CO-reduced CO dehydrogenase enzyme complex from Methanosarcina thermophila; the apparent Km and Vmax values were 1.7 +/- 0.3 mM TCE and 26.2 +/- 1.7 mol TCE dechlorinated/min/mmol factor III. Factor III also catalysed the dechlorination of TCE when in the presence of titanium(III) citrate; the apparent Km and Vmax values were 1.2 +/- 0.3 mM TCE and 34.9 +/- 3.6 mol TCE dechlorinated/min/mmol factor III. The enzyme complex was resolved into the two-subunit nickel/iron-sulfur (Ni/Fe-S) component and the two-subunit factor III-containing corrinoid/iron-sulfur (Co/Fe-S) component. The Ni/Fe-S component was unable to dechlorinate TCE in the presence of CO; however, reconstitution with the Co/Fe-S component yielded the same dechlorinated products as with the CO dehydrogenase enzyme complex.  相似文献   

17.
Dehydroepiandrosterone sulfate is the most abundant sulfated steroid transformed in human tissues and serves as a precursor for steroid hormones. Recombinant human dehydroepiandrosterone sulfotransferase (DHEA-ST) expressed in glutathione sulfotransferase fusion form in E. coli was purified using glutathione sepharose 4B affinity adsorption chromatography, a Factor Xa cleavage step, and Q-sepharose fast flow column chromatography. The homogeneous preparation had an activity toward dehydroepiandrosterone (DHEA) of 150+/-40 nmol/min per mg of protein under the assay conditions at an overall yield of 38.4%. The recombinant human DHEA-ST was shown to have a subunit mass of 34 kDa by sodium dodecyl sulfate polyacrylamide gel electrophoresis, while having a molecular mass of 67.2 kDa by Superose-12 gel filtration. Our results indicate that the active recombinant enzyme expressed in E. coli is a homodimer.Biochemical properties for purified DHEA-ST were studied using DHEA as a substrate. The optimum pH ranged from pH 7 to 8, and the optimum temperature 40-45 degrees C. Ninety percent of basal DHEA-ST activity remained even after the enzyme was treated at 45 degrees C for 15 min. The 50% inactivation concentration of NaCl for DHEA-ST activity was determined to be around 500 mM. The K(m) value for DHEA was 1.9+/-0.3 microM and V(max)=190+/-18 nmol/min per mg of protein at 37 degrees C, pH 7.5.  相似文献   

18.
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   

19.
The kinetic properties of the rat intestinal microsomal 1-naphthol:UDPglucuronosyltransferase (EC 2.4.1.17) were investigated in fully activated microsomes prepared from isolated mucosal cells. The enzyme appeared to follow an ordered sequential bireactant mechanism in which 1-naphthol and UDP-glucuronic acid (UDPGlcUA) are the first and second binding substrates and UDP and 1-naphthol glucuronide the first and second products, respectively. Bisubstrate kinetic analysis yielded the following kinetic constants: Vmax = 102 +/- 6 nmol/min per mg microsomal protein, Km (UDPGlcUA) = 1.26 +/- 0.10 mM, Km (1-naphthol) = 96 +/- 10 microM and Ki (1-naphthol) = 25 +/- 7 microM. The rapid equilibrium random or ordered bireactant mechanisms, as well as the iso-Theorell-Chance mechanism, could be excluded by endproduct inhibition studies with UDP.UDP-N-acetylglucosamine (UDPGlcNAc), usually found to be an activator of UDP glucuronosyltransferase in liver microsomes, acted as a full competitive inhibitor towards UDPGlcUA in rat intestinal microsomes. With regard to 1-naphthol UDPGlcNAc exhibited a dual effect: both inhibition and activation was observed. The effect of activation by MgCl2 and Triton X-100 on the kinetic constants and the inhibition patterns of UDP and UDPGlcNAc were investigated. The results obtained suggest that latency in rat intestinal microsomes may be due to endproduct inhibition by UDP. This endproduct inhibition could be abolished by in vitro treatment with MgCl2 and Triton X-100.  相似文献   

20.
Lipid A disaccharide synthase of Escherichia coli catalyzes the reaction 2,3-diacyl-GlcN-1-P + UDP-2,3-diacyl-GlcN----2',3'-diacyl-GlcN (beta,1'----6)2,3-diacyl-GlcN-1-P + UDP (Ray, B. L., Painter, G., and Raetz, C. R. H. (1984) J. Biol. Chem. 259, 4852-4859). Using a strain that overproduces the enzyme about 200-fold we have devised a simple purification to near homogeneity, utilizing two types of dye-ligand resins and heparin-agarose. The overall purification starting with membrane-free extracts was 54-fold (16,000-fold relative to wild-type extracts) with a 31% yield. The subunit molecular mass determined by sodium dodecyl sulfate gel electrophoresis is approximately 42,000 daltons, and the native enzyme appears to be a dimer. The amino-terminal sequence is (X)-(Thr)-Glu-Gln-(X)-Pro-Leu-Thr-Ie-Ala..., consistent with the results predicted from the DNA sequence, Met-Thr-Glu-Gln-Arg-Pro-Leu-Thr-Ile-Ala.... The purified enzyme displays a strong kinetic preference for sugar substrates bearing two fatty acyl moieties, but it is, nevertheless, very useful for the semisynthetic preparation of many lipid A analogs. Gel filtration studies demonstrate that the natural substrates (2,3-diacyl-GlcN-1-P and UDP-2,3-diacyl-GlcN) form micelles (n approximately equal to 300), rather than bilayers, under conditions used to assay the enzyme. Unlike most enzymes of glycerophospholipid synthesis, the lipid A disaccharide synthase does not require the presence of a detergent for catalytic activity. At 1 mM UDP-2,3-diacyl-GlcN the Vmax and Km values for 2,3-diacyl-GlcN-1-P are 14,028 +/- 513 nmol/min/mg and 0.27 +/- 0.02 mM. When 2,3-diacyl-GlcN-1-P is maintained at 1 mM, they are 12,368 +/- 472 nmol/min/mg and 0.11 +/- 0.01 mM for UDP-2,3-diacyl-GlcN.  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