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.
An enzyme capable of dehalogenating vicinal haloalcohols to their corresponding epoxides was purified from the 3-chloro-1,2-propanediol-utilizing bacterium Arthrobacter sp. strain AD2. The inducible haloalcohol dehalogenase converted 1,3-dichloro-2-propanol, 3-chloro-1,2-propanediol, 1-chloro-2-propanol, and their brominated analogs, 2-bromoethanol, as well as chloroacetone and 1,3-dichloroacetone. The enzyme possessed no activity for epichlorohydrin (3-chloro-1,2-epoxypropane) or 2,3-dichloro-1-propanol. The dehalogenase had a broad pH optimum at about 8.5 and a temperature optimum of 50 degrees C. The enzyme followed Michaelis-Menten kinetics, and the Km values for 1,3-dichloro-2-propanol and 3-chloro-1,2-propanediol were 8.5 and 48 mM, respectively. Chloroacetic acid was a competitive inhibitor, with a Ki of 0.50 mM. A subunit molecular mass of 29 kDa was determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. With gel filtration, a molecular mass of 69 kDa was found, indicating that the native protein is a dimer. The amino acid composition and N-terminal amino acid sequence are given.  相似文献   

2.
Haloalcohol dehalogenases are bacterial enzymes that catalyze the cofactor-independent dehalogenation of vicinal haloalcohols such as the genotoxic environmental pollutant 1,3-dichloro-2-propanol, thereby producing an epoxide, a chloride ion and a proton. Here we present X-ray structures of the haloalcohol dehalogenase HheC from Agrobacterium radiobacter AD1, and complexes of the enzyme with an epoxide product and chloride ion, and with a bound haloalcohol substrate mimic. These structures support a catalytic mechanism in which Tyr145 of a Ser-Tyr-Arg catalytic triad deprotonates the haloalcohol hydroxyl function to generate an intramolecular nucleophile that substitutes the vicinal halogen. Haloalcohol dehalogenases are related to the widespread family of NAD(P)H-dependent short-chain dehydrogenases/reductases (SDR family), which use a similar Ser-Tyr-Lys/Arg catalytic triad to catalyze reductive or oxidative conversions of various secondary alcohols and ketones. Our results reveal the first structural details of an SDR-related enzyme that catalyzes a substitutive dehalogenation reaction rather than a redox reaction, in which a halide-binding site is found at the location of the NAD(P)H binding site. Structure-based sequence analysis reveals that the various haloalcohol dehalogenases have likely originated from at least two different NAD-binding SDR precursors.  相似文献   

3.
Biotransformation of 1,3-dichloro-2-propanol (DCP) to epichlorohydrin (ECH) by the whole cells of recombinant Escherichia coli expressing halohydrin dehalogenase was limited by product inhibition. To solve this problem and improve the ECH yield, a biotransformation strategy using resin-based in situ product removal (ISPR) was investigated. Seven macroporous resins were examined to adsorb ECH: resin HZD-9 was the best. When 10 % (w/v) HZD-9 was added to batch biotransformation, 53.3 mM ECH was obtained with a molar yield of 88.3 %. The supplement of the HZD-9 increased the ECH volumetric productivity from 0.5 to 2.8 mmol/l min compared to without addition of resin. In fed-batch biotransformation, this approach increased ECH from 31 to 87 mM. These results provide a promising basis for the biosynthesis of ECH.  相似文献   

4.
Epoxide hydrolase from Aspergillus niger was immobilized onto the modified Eupergit C 250 L through a Schiff base formation. Eupergit C 250 L was treated with ethylenediamine to introduce primary amine groups which were subsequently activated with glutaraldehyde. The amount of introduced primary amine groups was 220 μmol/g of the support after ethylenediamine treatment, and 90% of these groups were activated with glutaraldehyde. Maximum immobilization of 80% was obtained with modified Eupergit C 250 L under the optimized conditions. The optimum pH was 7.0 for the free epoxide hydrolase and 6.5 for the immobilized epoxide hydrolase. The optimum temperature for both free and immobilized epoxide hydrolase was 40 °C. The free epoxide hydrolase retained 52 and 33% of its maximum activity at 40 and 60 °C, respectively after 24 h preincubation time whereas the retained activities of immobilized epoxide hydrolase at the same conditions were 90 and 75%, respectively. Immobilized epoxide hydrolase showed about 2.5-fold higher enantioselectivity than that of free epoxide hydrolase. A preparative-scale (120 g/L) kinetic resolution of racemic styrene oxide using immobilized preparation was performed in a batch reactor and (S)-styrene oxide and (R)-1-phenyl-1,2-ethanediol were both obtained with about 50% yield and 99% enantiomeric excess. The immobilized epoxide hydrolase was retained 90% of its initial activity after 5 reuses.  相似文献   

5.
The enantioselective hydrolysis of racemic epichlorohydrin for the production of enantiopure (S)-epichlorohydrin using whole cells of Aspergillus niger ZJB-09173 in organic solvents was investigated. Cyclohexane was used as the reaction medium based on the excellent enantioselectivity of epoxide hydrolase from A. niger ZJB- 09173 in cyclohexane. However, cyclohexane had a negative effect on the stability of epoxide hydrolase from A. niger ZJB-09173. In the cyclohexane medium, substrate inhibition, rather than product inhibition of catalysis, was observed in the hydrolysis of racemic epichlorohydrin using A. niger ZJB-09173. The racemic epichlorohydrin concentration was markedly increased by continuous feeding of substrate without significant decline of the yield. Ultimately, 18.5% of (S)-epichlorohydrin with 98 percent enantiomeric excess from 153.6 mM of racemic epichlorohydrin was obtained by the dry cells of A. niger ZJB-09173, which was the highest substrate concentration in the production of enantiopure (S)-epichlorohydrin by epoxide hydrolases using an organic solvent medium among the known reports.  相似文献   

6.
Kinetics of Bacterial Growth on Chlorinated Aliphatic Compounds   总被引:12,自引:8,他引:4       下载免费PDF全文
With the pure bacterial cultures Ancylobacter aquaticus AD20 and AD25, Xanthobacter autotrophicus GJ10, and Pseudomonas sp. strain AD1, Monod kinetics was observed during growth in chemostat cultures on 1,2-dichloroethane (AD20, AD25, and GJ10), 2-chloroethanol (AD20 and GJ10), and 1,3-dichloro-2-propanol (AD1). Both the Michaelis-Menten constants (Km) of the first catabolic (dehalogenating) enzyme and the Monod half-saturation constants (Ks) followed the order 2-chloroethanol, 1,3-dichloro-2-propanol, epichlorohydrin, and 1,2-dichloroethane. The Ks values of strains GJ10, AD20, and AD25 for 1,2-dichloroethane were 260, 222, and 24 μM, respectively. The low Ks value of strain AD25 was correlated with a higher haloalkane dehalogenase content of this bacterium. The growth rates of strains AD20 and GJ10 in continuous cultures on 1,2-dichloroethane were higher than the rates predicted from the kinetics of the haloalkane dehalogenase and the concentration of the enzyme in the cells. The results indicate that the efficiency of chlorinated compound removal is indeed influenced by the kinetic properties and cellular content of the first catabolic enzyme. The cell envelope did not seem to act as a barrier for permeation of 1,2-dichloroethane.  相似文献   

7.
Enzymes IIa and IIb, which catalyze the conversion of epichlorohydrin (ECH) to 3-chloro-1,2-propanediol (MCP), were purified from Corynebacterium sp. strain N-1074, which catalyzes the formation of (R)-MCP from prochiral 1,3-dichloro-2-propanol via ECH. The specific activity of enzyme IIa for the formation of MCP from ECH was about 6.4-fold higher than that of enzyme IIb. Both enzymes catalyzed the conversion of 1,2-epoxides to the corresponding diol, although they differed in several enzymatic properties.  相似文献   

8.
A gene encoding halohydrin dehalogenase (HHDH) from Agrobacterium tumefaciens CCTCC M 87071 was cloned and expressed in Escherichia coli. To increase activity and stability of HHDH, 14 amino acid residues around the active site and substrate-binding pocket based on the structural analysis and molecular docking were selected as targets for site-directed mutagenesis. The studies showed that the mutant HHDH (Mut-HHDH) enzyme had a more accessible substrate-binding pocket than the wild-type HHDH (Wt-HHDH). Molecular docking revealed that the distance between the substrate and active site was closer in mutant which improved the catalytic activity. The expressed Wt-HHDH and Mut-HHDH were purified and characterized using 1,3-dichloro-2-propanol (1,3-DCP) as substrates. The specific activity of the mutant was enhanced 26-fold and the value of k cat was 18.4-fold as compared to the Wt-HHDH, respectively. The Mut-HHDH showed threefold extension of half-life at 45 °C than that of Wt-HHDH. Therefore it is possible to add 1,3-DCP concentration up to 100 mM and epichlorohydrin (ECH) was produced at a relatively high conversion and yield (59.6 %) using Mut-HHDH as catalyst. This Mut-HHDH could be a potential candidate for the upscale production of ECH.  相似文献   

9.
The production of (R)-3-chloro-1,2-propanediol [(R)-MCP] from prochiral 1,3-dichloro-2-propanol (DCP) was examined with a bacterial strain identified as a Corynebacterium strain. The addition of glycerol as a carbon source or some chlorinated alcohols to a medium was effective for the induction of activity catalyzing the transformation of DCP into MCP. The optimum pH for (R)-MCP production by the resting cell reaction was around 8.0. The optical purity of (R)-MCP formed was improved by keeping the level of DCP in the reaction mixture at a low concentration. (R)-MCP was obtained from 77.5 mM DCP with a 97.3% molar conversion yield and an 83.8% enantiomeric excess of its optical purity by periodic feeding of the substrate.  相似文献   

10.
In this study, a novel strain of Pichia jadinii, HBY61, capable of the biocatalysis of 4-hydroxy-2-butanone (4H2B) to (R)-1,3-BD was isolated. HBY61 produced (R)-1,3-BD with high activity and absolute stereochemical selectivity (100 % e.e). Glucose and beef extract were found to be the key factors governing the fermentation, and their optimal concentrations were determined to be 84.2 and 43.7 g/L, respectively. The optimal bioconversion conditions of 4H2B catalyzed by HBY61 were pH 7.4, 30 °C, and 250 rpm with 6 % (v/v) glucose as the co-substrate. Accordingly, when 45 g/L of 4H2B was divided into three equal parts and added successively into the system at set time intervals, the maximum (R)-1,3-BD concentration reached 38.3 g/L with high yield (85.1 %) and strict 100 % enantioselectivity. Compared with previously reported yields for the biocatalytic production of (R)-1,3-BD, the use of strain HBY61 provided a high yield with excellent stereoselectivity.  相似文献   

11.
2,3-Dichloro-1-propanol is more chemically stable than its isomer, 1,3-dichloro-2-propanol, and is therefore more difficult to degrade. The isolation of bacteria capable of complete mineralization of 2,3-dichloro-1-propanol was successful only from enrichments at high pH. The bacteria thus isolated were found to be members of the α division of the Proteobacteria in the Rhizobium subdivision, most likely Agrobacterium sp. They could utilize both dihaloalcohol substrates and 2-chloropropionic acid. The growth of these strains in the presence of 2,3-dichloro-1-propanol was strongly affected by the pH and buffer strength of the medium. Under certain conditions, a ladder of four active dehalogenase bands could be visualized from this strain in activity gels. The enzyme involved in the complete mineralization of 2,3-dichloro-1-propanol was shown to have a native molecular weight of 114,000 and consisted of four subunits of similar molecular weights.  相似文献   

12.
1,2,3-Trichloropropane (TCP) is a toxic compound that is recalcitrant to biodegradation in the environment. Attempts to isolate TCP-degrading organisms using enrichment cultivation have failed. A potential biodegradation pathway starts with hydrolytic dehalogenation to 2,3-dichloro-1-propanol (DCP), followed by oxidative metabolism. To obtain a practically applicable TCP-degrading organism, we introduced an engineered haloalkane dehalogenase with improved TCP degradation activity into the DCP-degrading bacterium Pseudomonas putida MC4. For this purpose, the dehalogenase gene (dhaA31) was cloned behind the constitutive dhlA promoter and was introduced into the genome of strain MC4 using a transposon delivery system. The transposon-located antibiotic resistance marker was subsequently removed using a resolvase step. Growth of the resulting engineered bacterium, P. putida MC4-5222, on TCP was indeed observed, and all organic chlorine was released as chloride. A packed-bed reactor with immobilized cells of strain MC4-5222 degraded >95% of influent TCP (0.33 mM) under continuous-flow conditions, with stoichiometric release of inorganic chloride. The results demonstrate the successful use of a laboratory-evolved dehalogenase and genetic engineering to produce an effective, plasmid-free, and stable whole-cell biocatalyst for the aerobic bioremediation of a recalcitrant chlorinated hydrocarbon.  相似文献   

13.
A gene encoding a putative epoxide hydrolase (EHase) was identified by analyzing an open reading frame of the genome sequence of Novosphingobium aromaticivorans, retaining the conserved catalytic residues such as the catalytic triad (Asp177, Glu328, and His355) and the oxyanion hole. The enantioselective EHase gene (neh) was cloned, and the recombinant EHase could be purified to apparent homogeneity by one step of metal affinity chromatography and further characterized. The purified N. aromaticivorans enantioselective epoxide hydrolase (NEH) showed enantioselective hydrolysis toward styrene oxide, glycidyl phenyl ether, epoxybutane, and epichlorohydrin. The optimal EHase activity toward styrene oxide occurred at pH 6.5 and 45°C. The purified NEH could preferentially hydrolyze (R)-styrene oxide with enantiomeric excess of more than 99% and 11.7% yield after 20-min incubation at an optimal condition. The enantioselective hydrolysis of styrene oxide was also confirmed by the analysis of the vicinal diol, 1-phenyl-1,2-ethanediol. The hydrolyzing rates of the purified NEH toward epoxide substrates were not affected by as high as 100 mM racemic styrene oxide.  相似文献   

14.
The use of enantioselective hydrolysis for preparing chiral epichlorohydrins was investigated using recombinantPichia pastoris with the enantioselective epoxide hydrolase ofRhodotorula glutinis. The rate of the recombinant epoxide hydrolase-catalyzed enantioselective hydrolysis of racemic epichlorohydrins was enhanced by the addition of 5% (v/v) Tween 20. Enantiopure (R)-epichlorohydrins with an enantiopurity of 100%ee and a yield of 26% were obtained within 5 min from 50 mM racemates.  相似文献   

15.
Using a combined strategy of random mutagenesis of haloalkane dehalogenase and genetic engineering of a chloropropanol-utilizing bacterium, we constructed an organism that is capable of growth on 1,2,3-trichloropropane (TCP). This highly toxic and recalcitrant compound is a waste product generated from the manufacture of the industrial chemical epichlorohydrin. Attempts to select and enrich bacterial cultures that can degrade TCP from environmental samples have repeatedly been unsuccessful, prohibiting the development of a biological process for groundwater treatment. The critical step in the aerobic degradation of TCP is the initial dehalogenation to 2,3-dichloro-1-propanol. We used random mutagenesis and screening on eosin-methylene blue agar plates to improve the activity on TCP of the haloalkane dehalogenase from Rhodococcus sp. m15-3 (DhaA). A second-generation mutant containing two amino acid substitutions, Cys176Tyr and Tyr273Phe, was nearly eight times more efficient in dehalogenating TCP than wild-type dehalogenase. Molecular modeling of the mutant dehalogenase indicated that the Cys176Tyr mutation has a global effect on the active-site structure, allowing a more productive binding of TCP within the active site, which was further fine tuned by Tyr273Phe. The evolved haloalkane dehalogenase was expressed under control of a constitutive promoter in the 2,3-dichloro-1-propanol-utilizing bacterium Agrobacterium radiobacter AD1, and the resulting strain was able to utilize TCP as the sole carbon and energy source. These results demonstrated that directed evolution of a key catabolic enzyme and its subsequent recruitment by a suitable host organism can be used for the construction of bacteria for the degradation of a toxic and environmentally recalcitrant chemical.  相似文献   

16.
Summary Pseudomonas sp. OS-K-29 assimilated (R)-2,3-dichloro-1-propanol preferentially as the sole source of carbon. Isolation of optically pure (S)-2,3-dichloro-1-propanol with 100% enantiomer excess (e.e.) from the racemate was done based on this bacterial assimilation using immobilized-cells of OS-K-29 with calcium-alginate. The overall examination of the reactor involved 19 batches for 50 days without loss of its activity. Highly pure (R)-epichlorohydrin with 99.5% e.e. was prepared from the (S)-2,3-dichloro-1-propanol with treatment of aqueous NaOH. This new method is simple and useful for manufacturing optically active (S)-2,3-dichloro-1-propanol and (R)-epichlorohydrin.  相似文献   

17.
The degradation of low concentrations of 1,3-dichloro-2-propanol (1,3-DCP) and related halohydrins by whole cells and cell-free extracts of soil bacteria has been investigated. Three bacteria (strains A1, A2, A4), isolated from the same soil sample, were distinguished on the basis of cell morphology, growth kinetics and haloalcohol dehalogenase profiles. Strain A1, probably an Agrobacterium sp., dehalogenated 1,3-DCP with the highest specific activity (0.33 U mg protein−1) and also had the highest affinity for 1,3-DCP (K m, 0.1 mM). Non-growing cells of this bacterium dehalogenated low concentrations of 1,3-DCP with a first-order rate constant (k 1) of 1.13 h−1 . The presence of a non-dehalogenating bacterium, strain G1 (tentatively identified as Pseudomonas mesophilius), did not enhance the dehalogenation rate of low 1,3-DCP concentrations. However, the mixed-species consortium of strains A1 and G1 had greater stability than the mono-species culture at DCP concentrations above 1.0 gl−1. Received: 30 April 1996 / Received revision: 30 July 1996 / Accepted: 5 August 1996  相似文献   

18.
Bacillus sp. Z018, a novel strain producing epoxide hydrolase, was isolated from soil. The epoxide hydrolase catalyzed the stereospecific hydrolysis of (R)-phenyl glycidyl ether to generate (R)-3-phenoxy-1,2-propanediol. Epoxide hydrolase from Bacillus sp. Z018 was inducible, and (R)-phenyl glycidyl ether was able to act as an inducer. The fermentation conditions for epoxide hydrolase were 35°C, pH 7.5 with glucose and NH4Cl as the best carbon and nitrogen source, respectively. Under optimized conditions, the biotransformation yield of 45.8% and the enantiomeric excess of 96.3% were obtained for the product (R)-3-phenoxy-1,2-propanediol.  相似文献   

19.
In order to produce enantiomerically pure epoxides for the synthesis of value-added chemicals, a novel putative epoxide hydrolase (EH) sgeh was cloned and overexpressed in pET28a/Escherichia coli BL21(DE3). The 1047 bp sgeh gene was mined from Streptomyces griseus NBRC 13350 genome sequence. The recombinant hexahistidyl-tagged SGEH was purified (16.6-fold) by immobilized metal-affinity chromatography, with 90% yield as a homodimer of 100 kDa. The recombinant E. coli whole cells overexpressing SGEH could kinetically resolve racemic phenyl glycidyl ether (PGE) into (R)-PGE with 98% ee, 40% yield, and enantiomeric ratio (E) of 20. This was achieved under the optimized reaction conditions i.e. cell/substrate ratio of 20:1 (w/w) at pH 7.5 and 20?°C in 10% (v/v) dimethylformamide (DMF) in a 10 h reaction. 99% enantiopure (R)-PGE was obtained when the reaction time was prolonged to 12 h with a yield of 34%. In conclusion, an economically viable and environment friendly green process for the production of enantiopure (R)-PGE was developed by using wet cells of E. coli expressing recombinant SGEH.  相似文献   

20.
In epoxide hydrolase from Agrobacterium radiobacter (EchA), phenylalanine 108 flanks the nucleophilic aspartate and forms part of the substrate-binding pocket. The influence of mutations at this position on the activity and enantioselectivity of the enzyme was investigated. Screening for improved enantioselectivity towards para-nitrophenyl glycidyl ether (pNPGE) using spectrophotometric progress curve analysis yielded five different mutants with 3- to 7-fold improved enantioselectivity. The increase in enantioselectivity was in most cases the result of an enhanced catalytic efficiency toward the preferred enantiomer. Several mutations at position F108 resulted in a higher activity toward cis-disubstituted meso-epoxides, which were converted to a single product enantiomer. Mutant F108C converted cis-2,3-epoxybutane to (2R,3R)-2,3-butanediol of >99% ee with a 7-fold improved activity, and mutant F108A hydrolyzed cyclohexene oxide to (1R,2R)-1,2-cyclohexanediol of >99% ee with a more than 150-fold higher activity than wild-type enzyme. It is concluded that single amino acid substitutions in the active site of epoxide hydrolase can result in enzyme variants with catalytic properties that are suitable for preparative scale production of (S)-epoxides and chiral vicinal diols in high yield and with excellent ee.  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