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.
C Zipper  K Nickel  W Angst    H P Kohler 《Applied microbiology》1996,62(12):4318-4322
Sphingomonas herbicidovorans MH (previously designated Flavobacterium sp. strain MH) was able to utilize the chiral herbicide (RS)-2-(4-chloro-2-methylphenoxy)propionic acid (mecoprop) as the sole carbon and energy source. When strain MH was offered racemic mecoprop as the growth substrate, it could degrade both the (R) and the (S) enantiomer to completion, as shown by biomass formation, substrate consumption, and stoichiometric chloride release. However, the (S) enantiomer disappeared much faster from the culture medium than the (R) enantiomer. These results suggest the involvement of specific enzymes for the degradation of each enantiomer. This view was substantiated by the fact that resting cells of strain MH grown on (S)-mecoprop were able to degrade the (S) but not the (R) enantiomer of mecoprop. Accordingly, resting cells of strain MH grown on (R)-mecoprop preferentially metabolized the (R) enantiomer. Nevertheless, such cells could transform (S)-mecoprop at low rates. Oxygen uptake rates with resting cells confirmed the above view, as oxygen consumption was strongly dependent on the growth substrate. Cells grown on (R)-mecoprop showed oxygen uptake rates more than two times higher upon incubation with the (R) than upon incubation with the (S) enantiomer and vice versa.  相似文献   

2.
(R)- and (S)-dichlorprop/alpha-ketoglutarate dioxygenases (RdpA and SdpA) catalyze the oxidative cleavage of 2-(2,4-dichlorophenoxy)propanoic acid (dichlorprop) and 2-(4-chloro-2-methyl-phenoxy)propanoic acid (mecoprop) to form pyruvate plus the corresponding phenol concurrent with the conversion of alpha-ketoglutarate (alphaKG) to succinate plus CO2. RdpA and SdpA are strictly enantiospecific, converting only the (R) or the (S) enantiomer, respectively. Homology models were generated for both enzymes on the basis of the structure of the related enzyme TauD (PDB code 1OS7). Docking was used to predict the orientation of the appropriate mecoprop enantiomer in each protein, and the predictions were tested by characterizing the activities of site-directed variants of the enzymes. Mutant proteins that changed at residues predicted to interact with (R)- or (S)-mecoprop exhibited significantly reduced activity, often accompanied by increased Km values, consistent with roles for these residues in substrate binding. Four of the designed SdpA variants were (slightly) active with (R)-mecoprop. The results of the kinetic investigations are consistent with the identification of key interactions in the structural models and demonstrate that enantiospecificity is coordinated by the interactions of a number of residues in RdpA and SdpA. Most significantly, residues Phe171 in RdpA and Glu69 in SdpA apparently act by hindering the binding of the wrong enantiomer more than the correct one, as judged by the observed decreases in Km when these side chains are replaced by Ala.  相似文献   

3.
Alpha-ketoglutarate-dependent (R)-dichlorprop dioxygenase (RdpA) and alpha-ketoglutarate-dependent (S)-dichlorprop dioxygenase (SdpA), which are involved in the degradation of phenoxyalkanoic acid herbicides in Sphingomonas herbicidovorans MH, were expressed and purified as His6-tagged fusion proteins from Escherichia coli BL21(DE3)(pLysS). RdpA and SdpA belong to subgroup II of the alpha-ketoglutarate-dependent dioxygenases and share the specific motif HXDX(24)TX(131)HX(10)R. Amino acids His-111, Asp-113, and His-270 and amino acids His-102, Asp-104, and His 257 comprise the 2-His-1-carboxylate facial triads and were predicted to be involved in iron binding in RdpA and SdpA, respectively. RdpA exclusively transformed the (R) enantiomers of mecoprop [2-(4-chloro-2-methylphenoxy)propanoic acid] and dichlorprop [2-(2,4-dichlorophenoxy)propanoic acid], whereas SdpA was specific for the (S) enantiomers. The apparent Km values were 99 microM for (R)-mecoprop, 164 microM for (R)-dichlorprop, and 3 microM for alpha-ketoglutarate for RdpA and 132 microM for (S)-mecoprop, 495 microM for (S)-dichlorprop, and 20 microM for alpha-ketoglutarate for SdpA. Both enzymes had high apparent Km values for oxygen; these values were 159 microM for SdpA and >230 microM for RdpA, whose activity was linearly dependent on oxygen at the concentration range measured. Both enzymes had narrow cosubstrate specificity; only 2-oxoadipate was able to replace alpha-ketoglutarate, and the rates were substantially diminished. Ferrous iron was necessary for activity of the enzymes, and other divalent cations could not replace it. Although the results of growth experiments suggest that strain MH harbors a specific 2,4-dichlorophenoxyacetic acid-converting enzyme, tfdA-, tfdAalpha-, or cadAB-like genes were not discovered in a screening analysis in which heterologous hybridization and PCR were used.  相似文献   

4.
Whole cells of Desulfobulbus propionicus fermented [1-13C]ethanol to [2-13C] and [3-13C]propionate and [1-13C]-acetate, which indicates the involvement of a randomizing pathway in the formation of propionate. Cell-free extracts prepared from cells grown on lactate (without sulfate) contained high activities of methylmalonyl-CoA: pyruvate transacetylase, acetase kinase and reasonably high activities of NAD(P)-independent L(+)-lactate dehydrogenase NAD(P)-independent pyruvate dehydrogenase, phosphotransacetylase, acetate kinase and reasonably high activity of NAD(P)-independent L(+)-lactate dehydrogenase, fumarate reductase and succinate dehydrogenase. Cell-free extracts catalyzed the conversion of succinate to propionate in the presence of pyruvate, CoA and ATP and the oxaloacetate-dependent conversion of propionate to succinate. After growth on lactate or propionate in the presence of sulfate similar enzyme levels were found except for fumarate reductase which was considerably lower. Fermentative growth on lactate led to higher cytochrome b contents than growth with sulfate as electron acceptor.The labeling studies and the enzyme measurements demonstrate that in Desulfobulbus propionate is formed via a succinate pathway involving a transcarboxylase like in Propionibacterium. The same pathway may be used for the degradation of propionate to acetate in the presence of sulfate.Abbreviations DCPIP 2,6-dichlorophenolindophenol - PEP phosphoenolpyruvate  相似文献   

5.
Cell extracts of Pseudomonas aeruginosa strain PAO were found to contain pyruvate carboxylase activity. Specific activities were minimal when cells were grown on Casamino Acids, acetate, or succinate, but were three- to fourfold higher when cells were grown in glucose, gluconate, glycerol, lactate, or pyruvate minimal media. The reaction in crude cell extracts and in partially purified preparations was dependent on pyruvate, adenosine 5'-triphosphate, and Mg(2+), but was not affected by either the presence or absence of acetyl coenzyme A. Activity was nearly totally inhibited by avidin and this inhibition was substantially blocked by free biotin in incubation mixtures. Cell extracts were shown to fix (14)CO(2) in a reaction that had these same characteristics. Eight pleiotropic, carbohydrate-negative mutant strains of the organism were isolated after nitrosoguanidine mutagenesis. Each mutant strain grew normally in acetate, succinate, and citrate minimal media but failed to utilize glucose, gluconate, 2-ketogluconate, mannitol, glycerol, lactate, and pyruvate as sole sources of carbon and energy. These strains were found by quantitative transductional analysis with phage F116 to form a single linkage group. Cell extracts of each mutant strain were either lacking or severely deficient in pyruvate carboxylase activity. Spontaneous revertants of five of the eight strains were isolated and found to recover simultaneously both pyruvate carboxylase activity and the ability to utilize each of the C(6) and C(3) compounds. A second linkage group of similar mutant strains that grew on the C(3) compounds was found to contain normal levels of pyruvate carboxylase activity, but each strain was deficient in an enzyme of the Entner-Doudoroff pathway.  相似文献   

6.
1. Cell-free extracts, prepared from a non-fluorescent Pseudomonas grown on m-cresol, oxidized gentisate and certain alkyl-substituted gentisates with the consumption of 1 mol of oxygen and the formation of 1 mol of pyruvate from 1 mol of substrate. 2. In addition to pyruvate, malate was formed from gentisate; citramalate was formed from 3-methylgentisate and 4-methylgentisate; 2,3-dimethylmalate was formed from 3,4-dimethylgentisate. 3. One enantiomer, d-(-)-citramalate, was formed enzymically from 3-methylgentisate, 4-methylgentisate and citraconate. l-(+)-Citramalate was formed from mesaconate by the same extracts. When examined as its dimethyl ester by gas-liquid chromatography, enzymically formed 2,3-dimethylmalate showed the same behaviour as one of the two racemates prepared from the synthetic compound. 4. Maleate, citraconate and 2,3-dimethylmaleate were rapidly hydrated by cell extracts, but ethylfumarate and 2,3-dimethylfumarate were not attacked. 5. Cell extracts oxidized 1,4-dihydroxy-2-naphthoate to give pyruvate and phthalate. 6. Alkylgentisates were oxidized by a gentisate oxygenase (EC 1.13.1.4) present in Pseudomonas 2,5. The ring-fission products were attacked by maleylpyruvase, but not by fumarylpyruvase, and their u.v.-absorption spectra were those expected for alkyl-substituted maleylpyruvates. 7. When supplemented with ATP, CoA, succinate and Mg(2+) ions, an enzyme system from cells grown with 2,5-xylenol formed pyruvate from d- but not from l-citramalate. Extracts from cells grown with dl-citramalate or with itaconate attacked both d- and l-citramalate; other alkylmalates were cleaved in similar fashion to give pyruvate or 2-oxobutyrate. 8. These results accord with a general sequence of reactions in which the benzene nucleus of an alkylgentisate is cleaved to give an alkyl-substituted maleylpyruvate. The ring-fission products are hydrolysed to give pyruvate, plus alkylmalic acids which then undergo aldol fissions, probably as their CoA esters. In Pseudomonas 2,5 several homologous sequences of this general type appear to be catalysed by a single battery of enzymes with broad substrate specificities, whereas the metabolic capabilities of the fluorescent Pseudomonas 3,5 are more restricted. 9. Intact cells of both organisms metabolize d-malic acid by reactions that have not been elucidated, but are different from those which degrade alkylmalates.  相似文献   

7.
On the basis of enzyme activities detected in extracts of Selenomonas ruminantium HD4 grown in glucose-limited continuous culture, at a slow (0.11 h-1) and a fast (0.52 h-1) dilution rate, a pathway of glucose catabolism to lactate, acetate, succinate, and propionate was constructed. Glucose was catabolized to phosphoenol pyruvate (PEP) via the Emden-Meyerhoff-Parnas pathway. PEP was converted to either pyruvate (via pyruvate kinase) or oxalacetate (via PEP carboxykinase). Pyruvate was reduced to L-lactate via a NAD-dependent lactate dehydrogenase or oxidatively decarboxylated to acetyl coenzyme A (acetyl-CoA) and CO2 by pyruvate:ferredoxin oxidoreductase. Acetyl-CoA was apparently converted in a single enzymatic step to acetate and CoA, with concomitant formation of 1 molecule of ATP; since acetyl-phosphate was not an intermediate, the enzyme catalyzing this reaction was identified as acetate thiokinase. Oxalacetate was converted to succinate via the activities of malate dehydrogenase, fumarase and a membrane-bound fumarate reductase. Succinate was then excreted or decarboxylated to propionate via a membrane-bound methylmalonyl-CoA decarboxylase. Pyruvate kinase was inhibited by Pi and activated by fructose 1,6-bisphosphate. PEP carboxykinase activity was found to be 0.054 mumol min-1 mg of protein-1 at a dilution rate of 0.11 h-1 but could not be detected in extracts of cells grown at a dilution rate of 0.52 h-1. Several potential sites for energy conservation exist in S. ruminantium HD4, including pyruvate kinase, acetate thiokinase, PEP carboxykinase, fumarate reductase, and methylmalonyl-CoA decarboxylase. Possession of these five sites for energy conservation may explain the high yields reported here (56 to 78 mg of cells [dry weight] mol of glucose-1) for S. ruminantium HD4 grown in glucose-limited continuous culture.  相似文献   

8.
Pyruvate carboxylase has been detected in, and partially purified from, cell-free extracts of Azotobacter vinelandii OP. The best preparations obtained have specific activities in the range of 4 units/mg and appear approximately 15% pure when analyzed by polyacrylamide gel electrophoresis. The partially purified enzyme is activated by both univalent and divalent cations, contains one or more functional biotinyl residues, and exhibits apparent Michaelis constants for the substrates (pyruvate, Mg-ATP2?, and HCO3?) which are in the same range as those observed for other pyruvate carboxylases. However, A. vinelandii pyruvate carboxylase is fully active in the absence of added acetyl-coenzyme A and is insensitive to inhibition by dicarboxylic acids such as l-aspartate, l-glutamate, and α-ketoglutarate. The molecular weight of the catalytically active species is obtained as 296,000.The level of pyruvate carboxylase is highest in extracts of A. vinelandii grown on pyruvate or l-lactate as sole carbon source and this level is further enhanced on addition of succinate to the medium. The enzyme is absent from cells grown on succinate and is present at intermediate levels in cells grown on sucrose, glucose, glycerol, or acetate. In contrast, the level of phosphoenolypyruvate carboxylase in these extracts is essentially independent of the carbon source. These data suggest that pyruvate carboxylase in A. vinelandii is induced by pyruvate or some closely related metabolite.  相似文献   

9.
Rhodococcus sp. strain DTB (DSM 44534) was grown on a mixture of (R,R)-, (S,S)- and meso-bis-(1-chloro-2-propyl) ether (BCPE) as the sole source of carbon and energy. During BCPE degradation 1'-chloro-2'-propyl-3-chloro-2-prop-1-enyl-ether (DVE), 1-chloro-2-propanol and chloroacetone intermediates were formed. The BCPE or DVE stereoisomers were metabolized in consecutive order via scission of the ether bond, with discrimination against the (R) configuration. Resting cell suspensions of Rhodococcus pregrown on BCPE showed a preferential attack of the (S)-configured ether-linked carbons, resulting in an enantioselective enrichment of (R,R)-BCPE. Microbial discrimination of BCPE or DVE isomers and chemical conversion of the intermediates to 1-chloro-2-propanol allowed the identification of the configuration of all BCPE isomers and the DVE enantiomers. Elucidation of the absolute configuration of the 1-chloro-2-propanol isomers was achieved by enantioselective chemical synthesis.  相似文献   

10.
Sphingomonas herbicidovorans MH was able to completely degrade both enantiomers of the chiral herbicide dichlorprop [(RS)-2-(2,4-dichlorophenoxy)propanoic acid], with preferential degradation of the (S) enantiomer over the (R) enantiomer. These results are in agreement with the recently reported enantioselective degradation of mecoprop [(RS)-2-(4-chloro-2-methylphenoxy)propanoic acid] by this bacterium (C. Zipper, K. Nickel, W. Angst, and H.-P. E. Kohler, Appl. Environ. Microbiol. 62:4318–4322, 1996). Uptake of (R)-dichlorprop, (S)-dichlorprop, and 2,4-D (2,4-dichlorophenoxyacetic acid) was inducible. Initial uptake rates of cells grown on the respective substrate showed substrate saturation kinetics with apparent affinity constants (Kt) of 108, 93, and 117 μM and maximal velocities (Vmax) of 19, 10, and 21 nmol min−1 mg of protein−1 for (R)-dichlorprop, (S)-dichlorprop, and 2,4-D, respectively. Transport of (R)-dichlorprop, (S)-dichlorprop, and 2,4-D was completely inhibited by various uncouplers and by nigericin but was only marginally inhibited by valinomycin and by the ATPase inhibitor N,N′-dicyclohexylcarbodiimine. Experiments on the substrate specificity of the putative transport systems revealed that (R)-dichlorprop uptake was inhibited by (R)-mecoprop but not by (S)-mecoprop, (S)-dichlorprop, or 2,4-D. On the other hand, the (S)-dichlorprop transport was inhibited by (S)-mecoprop but not by (R)-mecoprop, (R)-dichlorprop, or 2,4-D. These results provide evidence that the first step in the degradation of dichlorprop, mecoprop, and 2,4-D by S. herbicidovorans is active transport and that three inducible, proton gradient-driven uptake systems exist: one for (R)-dichlorprop and (R)-mecoprop, another for (S)-dichlorprop and (S)-mecoprop, and a third for 2,4-D.  相似文献   

11.
Desulfitobacterium chlororespirans Co23 is capable of using 3-chloro-4-hydroxybenzoate as terminal electron acceptor for growth. Membrane preparations from cells grown fermentatively on pyruvate in the presence of 3-chloro-4-hydroxybenzoate dechlorinated this compound at a rate of 3.9 nmol min(sup-1) mg of protein(sup-1). Fivefold-greater dechlorination rates were measured with reduced methyl viologen as the artificial electron donor. Reduced benzyl viologen, NADH, NADPH, reduced flavin adenine dinucleotide, and reduced flavin mononucleotide could not substitute for reduced methyl viologen. The maximal initial rate of catalysis was achieved at pH 6.5 and 60(deg)C. The membrane-bound dechlorinating enzyme system was not oxygen sensitive and was stable at 57(deg)C for at least 2 h. Sulfite inhibited dechlorination in cell-free assays, whereas sulfate did not. Several chlorophenols were dehalogenated exclusively in the ortho position by cell extracts.  相似文献   

12.
The fumarate reductase (NADH) present in cell-free extracts of S. lactis C10 was purified approximately 100-fed by chromatography on DEAE-cellulose in the presence of the non-ionic detergent Teric X-10, and some of the properties of this partially purified enzyme were characterized. Fumarate was able to act as a terminal electron acceptor and decreased the amount of lactate formed and oxygen used during the metabolism of pyruvate by resting cells of S. lactis. Anaerobic growth of S. lactis on glycerol was not observed and fumarate reduction was not coupled with glycerol-3-phosphate oxidation.  相似文献   

13.
Epoxide metabolism in Xanthobacter autotrophicus Py2 results in the conversion of epoxypropane to acetoacetate. Epoxide metabolism is initiated by the nucleophilic addition of coenzyme M to the (R)- and (S)-enantiomers of epoxypropane which forms the respective enantiomers of 2-hydroxypropyl-coenyme M. The (R)- and (S)-enantiomers of 2-hydroxypropyl coenzyme are oxidized to the achiral product 2-ketopropyl-CoM by two stereoselective dehydrogenases. The dehydrogenases catalyzing these reactions, termed (R)-hydroxypropyl-coenzyme M dehydrogenase (R-HPCDH) and (S)-hydroxypropyl-coenzyme M dehydrogenase (S-HPCDH), are NAD(+)-dependent enzymes belonging to the short chain dehydrogenase/reductase (SDR) family of enzymes. In this study, the crystal structure of R-HPCDH cocrystallized in the presence of (S)-hydroxypropyl-coenzyme M has been determined using X-ray diffraction methods and refined to 1.8 A resolution. The structure of R-HPCDH is tetrameric and stabilized by the interaction of the terminal carboxylates of each subunit with divalent metal ions. The structure of the presumed product-bound state reveals that binding interactions between the negatively charged oxygen atoms of the sulfonate moiety have striking similarities to sulfonate interactions observed in the previously determined structure of 2-ketopropyl-CoM oxidoreductase/carboxylase, highlighting the utility of coenzyme M as a carrier molecule in the pathway. The key elements of the aforementioned interactions are electrostatic interactions between the sulfonate oxygen atoms and two arginine residues (R152 and R196) of R-HPCDH. The comparison of the structure of R-HPCDH with a homology model of S-HPCDH provides a structural basis for a mechanism of substrate specificity in which the binding of the substrate sulfonate moiety at distinct sites on each stereoselective enzyme directs the orientation of the appropriate substrate enantiomer for hydride abstraction.  相似文献   

14.
Purple nonsulfur bacteria, Rhodospirillum rubrum and Rhodopseudomonas spheroides were found to possess coenzyme B12-dependent glutamate mutase activity. Cell-free extracts of these bacteria grown on Co2+-containing media catalyzed the conversion of glutamate to β-methylaspartate and further to mesaconate. The activity of the cell-free extracts of these organisms cultivated on Co2+-deficient media was markedly lower than that of the normal cells. Addition of coenzyme B12 to the former reaction mixture enhanced the mesaconate formation via β-methylaspartate. These results indicate the involvement of coenzyme Independent glutamate mutase of these bacteria in the dissimilation of glutamate to acetyl-CoA and pyruvate through the following pathway.

glutamate→β→methylaspartate→mesaconate→citramalate→→acetyl-CoA, pyruvate On the other hand, a greater part of glutamate was converted to α-hydroxyglutarate and succinate with the cell-free extracts of these photosynthetic bacteria. This fact, taking account of the presence of propionyl-CoA carboxylase in these bacteria, implies the participation of coenzyme B12-dependent (R)-methylmalonyl-CoA mutase in the formation of succinate via the following route.

glutamate→α-ketoglutarate→α-hydroxyglutarate→propionate→propionyl-CoA→(S)-methylmalonyl-CoA→(R)-methylmalonyl-CoA→succinyl-CoA  相似文献   

15.
Saccharomyces cerevisiae (baker's yeast) reduces ethyl 4-chloro-acetoacetate enantioselectively to ( R )- or ( S )-ethyl 4-chloro-3-hydroxybutyrate depending on the reaction conditions and the physiological state of the yeast cells. The ( S )-enantiomer is obtained under batch conditions with resting cells (55%, enantiomeric excess [ee]), and 4-chloro-acetate fed-batch actively metabolising yeast affords the ( R )-isomer (54%, ee). The enantioselective reduction of the substrate is accompanied by competing enzyme actions. Of the metabolites formed from the substrate, chloroacetone and the target compound ( R )-ethyl 4-chloro-3-hydroxybutyrate emerged as most important effectors of enantioselectivity of the microbial reduction. As a minor side-reaction, an aerobic reductive dehalogenation of the substrate was observed. The unusual high enantiopurity of the dehalo-product ( S )-ethyl 3-hydroxybutyrate confirms the stereodirecting effect of chloroacetone impressively. Hence, with S. cerevisiae either enantiomer can be obtained by variation of reaction conditions. The yeast further turned out to be a promising biocatalyst for dehalogenations.  相似文献   

16.
Carbon and ammonia metabolism of Spirillum lipoferum.   总被引:14,自引:11,他引:3       下载免费PDF全文
Intact cells and extracts from Spirillum lipoferum rapidly oxidized malate, succinate, lactate, and pyruvate. Glucose, galactose, fructose, acetate, and citrate did not increase the rate of O2 uptake by cells above the endogenous rate. Cells grown on NH+/4 oxidized the various substrates at about the same rate as did cells grown on N2. Added oxidized nicotinamide adenine dinucleotide generally enhanced O2 uptake by extracts supplied organic acids, whereas oxidized nicotinamide adenine dinucleotide phosphate had little effect. Nitrogenase synthesis repressed by growth of cells in the presence of NH+/4 was derepressed by methionine sulfoximine or methionine sulfone. The total glutamine synthetase activity from N2-grown cells was about eight times that from NH+/4-grown S. lipoferum; the response of glutamate dehydrogenase was the opposite. The total glutamate synthetase activity from N2-grown S. lipoferum was 1.4 to 2.6 times that from NH+/4-grown cells. The levels of poly-beta-hydroxybutyrate and beta-hydroxybutyrate dehydrogenase were elevated in cells grown on N2 as compared with those grown on NH+/4. Cell-free extracts capable of reducing C2H2 have been prepared; both Mg2+ and Mn2+ are required for good activity.  相似文献   

17.
研究了在好氧培养基中分别添加不同碳源对两阶段发酵菌体生长、酶活及代谢产物分布的影响,结果表明添加4mmol/L葡萄糖和12,54,80mmol/L乙酸钠均可以提高好氧阶段的菌体密度和相关酶活。将不同条件下培养的菌体转接厌氧发酵后,厌氧阶段的酶活和代谢产物分布也发生改变。进一步对酶活及代谢产物分析表明:Escherichia coli NZN111(sfcA)厌氧发酵过程中,磷酸烯醇式丙酮酸羧化激酶(PCK)是产丁二酸的关键酶,丙酮酸激酶(PYK)主要和副产物丙酮酸的积累有关,异柠檬酸裂解酶(ICL)对丁二酸产量也有一定影响。好氧培养基中添加80mmol/L乙酸钠,厌氧发酵结束时丁二酸的质量收率可达89.0%,相比对照提高了16.6%。  相似文献   

18.
Saccharomyces cerevisiae (baker's yeast) reduces ethyl 4-chloro-acetoacetate enantioselectively to ( R )- or ( S )-ethyl 4-chloro-3-hydroxybutyrate depending on the reaction conditions and the physiological state of the yeast cells. The ( S )-enantiomer is obtained under batch conditions with resting cells (55%, enantiomeric excess [ee]), and 4-chloro-acetate fed-batch actively metabolising yeast affords the ( R )-isomer (54%, ee). The enantioselective reduction of the substrate is accompanied by competing enzyme actions. Of the metabolites formed from the substrate, chloroacetone and the target compound ( R )-ethyl 4-chloro-3-hydroxybutyrate emerged as most important effectors of enantioselectivity of the microbial reduction. As a minor side-reaction, an aerobic reductive dehalogenation of the substrate was observed. The unusual high enantiopurity of the dehalo-product ( S )-ethyl 3-hydroxybutyrate confirms the stereodirecting effect of chloroacetone impressively. Hence, with S. cerevisiae either enantiomer can be obtained by variation of reaction conditions. The yeast further turned out to be a promising biocatalyst for dehalogenations.  相似文献   

19.
Six yeasts were studied for their ability to reduce ethyl 4-chloroacetoacetate (ethyl 4-chloro-3-oxobutanoate) stereoselectively. Five species reduced the substrate to ethyl (S)-4-chloro-3-hydroxybutanoate of high (92-99%) optical purity. With glucose-grown cells, substrate reduction could only be demonstrated when growth was oxygen-limited, whereas xylose-grown Pichia capsulata could be grown under conditions of oxygen excess without losing its reducing ability. Zygosaccha-romyces rouxii exhibited high enantioselectivity (≥98% ee (S)-enantiomer) under all conditions tested, whilst in P. capsulata, a novel switch was observed from producing mainly the (R)-enantiomer using glucose as co-substrate to producing mainly the (R)-enantiomer using 2-propanol as co-substrate. This switch was correlated with a change in reduction predominantly from an NADPH-dependent dehydrogenase system to an NADH-dependent system. In the production of ethyl (R)-4-chloro-3-hydroxybutanoate with P. capsulata, the enantioselectivity was also found to depend upon growth conditions. With glucose-grown cells, higher enantioselectivity was observed using cells harvested in stationary phase (93-94% ee) compared with cells harvested in exponential phase (43-60% ee). Growing P. capsulata with xylose rather than glucose as the major source of carbon for growth resulted in an eight-fold increase in the specific rate of ethyl (R)-4-chloro-3-hydroxybutanoate production using 2-propanol as co-substrate, although enantioselectivity was slightly reduced (65-81% ee) compared with the maximum achieved with glucose-grown cells. The effect of growth on xylose could also be correlated with enhanced activity of an NADH-dependent (R)-selective dehydrogenase system.  相似文献   

20.
When 3,4-dihydroxyphenylacetic acid (homoprotocatechuic acid) was added to Pseudomonase acidovorans growing at the expense of succinate, enzymes required for degrading homoprotocatechuate to pyruvate and succinate semialdehyde were strongly induced. These enzymes were effectively absent from cell extracts of the organism grown with 4-hydroxyphenylacetic acid, and this substrate was metabolized by the catabolic enzymes of the homogentisate pathway. Two separate ring-fission dioxygenases for 3,4,5-trihydroxybenzoic acid (gallic acid) were present in cell extracts of Pseudomonas putida when grown with syringic acid, and gallate was degraded by reactions associated with meta fission. One of the two gallate dioxygenases also attacked 3-O-methylgallic acid; the other, which did not, was induced when cells were exposed to gallate. This organism possessed ortho fission enzymes, including protocatechuate 3,4-dioxygenase (EC 1.13.11.3) and cis,cis-carboxymuconate-lactonizing enzyme (EC 5.5.1.2), after induction with 3,4-dihydroxybenzoic acid (protocatechuic acid). Gallate was a substrate for protocatechuate 3,4-dioxygenase, with a Vmax about 3% of that of protocatechuate and with an apparent Km slightly lower. Gallate was a powerful competitive inhibitor of protocatechuate oxidation.  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