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 62 毫秒
1.
Mesaconase catalyzes the hydration of mesaconate (methylfumarate) to (S)-citramalate. The enzyme participates in the methylaspartate pathway of glutamate fermentation as well as in the metabolism of various C5-dicarboxylic acids such as mesaconate or L-threo-β-methylmalate. We have recently shown that Burkholderia xenovorans uses a promiscuous class I fumarase to catalyze this reaction in the course of mesaconate utilization. Here we show that classical Escherichia coli class I fumarases A and B (FumA and FumB) are capable of hydrating mesaconate with 4% (FumA) and 19% (FumB) of the catalytic efficiency k cat/K m, compared to the physiological substrate fumarate. Furthermore, the genomes of 14.8% of sequenced Enterobacteriaceae (26.5% of E. coli, 90.6% of E. coli O157:H7 strains) possess an additional class I fumarase homologue which we designated as fumarase D (FumD). All these organisms are (opportunistic) pathogens. fumD is clustered with the key genes for two enzymes of the methylaspartate pathway of glutamate fermentation, glutamate mutase and methylaspartate ammonia lyase, converting glutamate to mesaconate. Heterologously produced FumD was a promiscuous mesaconase/fumarase with a 2- to 3-fold preference for mesaconate over fumarate. Therefore, these bacteria have the genetic potential to convert glutamate to (S)-citramalate, but the further fate of citramalate is still unclear. Our bioinformatic analysis identified several other putative mesaconase genes and revealed that mesaconases probably evolved several times from various class I fumarases independently. Most, if not all iron-dependent fumarases, are capable to catalyze mesaconate hydration.  相似文献   

2.
Two strains of S-citramalate-fermenting strictly anaerobic non-spore-formers were isolated in pure culture from anoxic mud samples of a creek and from a pond. One of them (strain CreCit 1) was studied in detail. It stained gram-negative, and contained β-hydroxymyristic acid. Nitrate, sulfate and other sulfur compounds were not utilized as electron acceptors. S-citramalate, citrate, mesaconate, and pyruvate were utilized as substrates; but R-citramalate, citraconate, l-glutamate, and carbohydrates not. S-citramalate was fermented to acetate, formate, and hydrogen. Citrate, mesaconate, and pyruvate were fermented to acetate and formate. The DNA base ratio was 59 mol% guanine plus cytosine. Strain CreCit 1 is described as a member of a new genus and a new species in the family Bacteroidaceae, Formivibrio citricus gen. nov., sp. nov.  相似文献   

3.
Rhodospirillum rubrum is among the bacteria that can assimilate acetate in the absence of isocitrate lyase, the key enzyme of glyoxylate shunt. Previously we have suggested the functioning of a new anaplerotic cycle of acetate assimilation in this bacterium: citramalate cycle, where acetyl-CoA is oxidized to glyoxylate. This work has demonstrated the presence of all the key enzymes of this cycle in R. rubrum extracts: citramalate synthase catalyzing condensation of acetyl-CoA and pyruvate with the formation of citramalate, mesaconase forming mesaconate from L-citramalate, and the enzymes catalyzing transformation of propionyl-CoA + glyoxylate 3-methylmalyl-CoA ? mesaconyl-CoA. At the same time, R. rubrum synthesizes crotonyl-CoA carboxylase/reductase, which is the key enzyme of ethylmalonyl-CoA pathway discovered recently in Rhodobacter sphaeroides. Physiological differences between the citramalate cycle and the ethylmalonyl-CoA pathway are discussed.  相似文献   

4.
Summary The oxidation of branched C5-dicarboxylates has been studied with microorganisms obtained from enrichement cultures with each of these acids. Bacteria grown on itaconate, mesaconate or methylsuccinate oxidize most of these compounds rapidly; (+) citramalate is oxidized 2–4 times as rapidly as (-) citramalate; and citraconate only feebly. Citraconate-grown cells do not oxidize itaconate and oxidize feebly mesaconate and methylsuccinate; they oxidize (-) citramalate 3–4 times faster than (+) citramalate. Cell-free extracts of various strains grown on different C5-dicarboxylates contain pronounced mesaconase activity, varying amounts of citraconase and hardly any itaconase activity.Communication number 633 from National Chemical Laboratory.  相似文献   

5.
The enzyme 3-methylaspartase (3-methylaspartate ammonia-lyase, EC 4.3.1.2) was found in the cells of enteric bacteria, especially in the genera Citrobacter and Morganella, that were grown under anoxic and oxygen-limited conditions. The enzymes were purified to homogeneity from the cell-free extracts of 18 active strains and had similar enzymological properties such as action on columns, specific activity, molecular weight, subunit structure, and N-terminal amino acid sequence similarity. The production of the enzyme was dependent on the limitation of oxygen during growth and was arrested by aeration. The addition of external electron acceptors such as dimethylsulfoxide could support cell growth and production of the enzyme. Activities of glutamate mutase (EC 5.4.99.1) and (S)-citramalate hydrolyase (EC 4.2.1.34), key enzymes of the mesaconate pathway of (S)-glutamate fermentation in the genus Clostridium, were detected in the cells of the active strains grown under oxygen-limited conditions. Based on the results, the mesaconate pathway is proposed to explain the (S)-glutamate fermentation process observed in Enterobacteriaceae, and 3-methylaspartase could be a marker enzyme for this pathway. Received: 28 May 1997 / Accepted: 16 July 1997  相似文献   

6.
Effects of Fe2+ ions on the levels of two enzymes (fumarase and mesaconase) with fumarase activity in two Pseudomonads grown under various nutritional conditions were investigated. Fe2+ ions decreased fumarase but increased mesaconase. A high level of mesaconase was found in Ps. arvilla which was unable to metabolize itaconate. The level of mesaconase in the itaconate-grown cells of Ps. fluorescens was almost the same as that in the glucose-grown cells. This suggests that mesaconase is not an enzyme involved in the metabolism of C5-branched-chain dicarboxylates but presumably, taking the place of fumarase, plays a role in the operation of the tricarboxylic acid cycle in the cells grown in the medium containing Fe2+ ions more than 10 nmol/ml.  相似文献   

7.
The mechanism of the aerobic dark assimilation of acetate in the photoheterotrophically grown purple nonsulfur bacteriumRhodospirillum rubrum was studied. Both in the light and in the dark, acetate assimilation inRsp. rubrum cells, which lack the glyoxylate pathway, was accompanied by the excretion of glyoxylate into the growth medium. The assimilation of propionate was accompanied by the excretion of pyruvate. Acetate assimilation was found to be stimulated by bicarbonate, pyruvate, the C4-dicarboxylic acids of the Krebs cycle, and glyoxylate, but not by propionate. These data implied that the citramalate (CM) cycle inRsp. rubrum cells can function as an anaplerotic pathway under aerobic dark conditions. This supposition was confirmed by respiration measurements. The respiration of cells oxidizing acetate depended on the presence of CO2 in the medium. The fact that the intermediates of the CM cycle (citramalate and mesaconate) markedly inhibited acetate assimilation but had almost no effect on cell respiration indicated that citramalate and mesaconate were intermediates of the acetate assimilation pathway. The inhibition of acetate assimilation and cell respiration by itaconate was due to its inhibitory effect on propionyl-CoA carboxylase, an enzyme of the CM cycle. The addition of 5 mM itaconate to extracts ofRsp. rubrum cells inhibited the activity of this enzyme by 85%. The data obtained suggest that the CM cycle continues to function inRsp. rubrum cells that have been grown anaerobically in the light and then transferred to the dark and incubated aerobically.  相似文献   

8.
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   

9.
In this work, we have compared the ability of Pandoraea pnomenusa B356 and of Burkholderia xenovorans LB400 to metabolize diphenylmethane and benzophenone, two biphenyl analogs in which the phenyl rings are bonded to a single carbon. Both chemicals are of environmental concern. P. pnomenusa B356 grew well on diphenylmethane. On the basis of growth kinetics analyses, diphenylmethane and biphenyl were shown to induce the same catabolic pathway. The profile of metabolites produced during growth of strain B356 on diphenylmethane was the same as the one produced by isolated enzymes of the biphenyl catabolic pathway acting individually or in coupled reactions. The biphenyl dioxygenase oxidizes diphenylmethane to 3-benzylcyclohexa-3,5-diene-1,2-diol very efficiently, and ultimately this metabolite is transformed to phenylacetic acid, which is further metabolized by a lower pathway. Strain B356 was also able to cometabolize benzophenone through its biphenyl pathway, although in this case, this substrate was unable to induce the biphenyl catabolic pathway and the degradation was incomplete, with accumulation of 2-hydroxy-6,7-dioxo-7-phenylheptanoic acid. Unlike strain B356, B. xenovorans LB400 did not grow on diphenylmethane. Its biphenyl pathway enzymes metabolized diphenylmethane, but they poorly metabolize benzophenone. The fact that the biphenyl catabolic pathway of strain B356 metabolized diphenylmethane and benzophenone more efficiently than that of strain LB400 brings us to postulate that in strain B356, this pathway evolved divergently to serve other functions not related to biphenyl degradation.  相似文献   

10.
《Insect Biochemistry》1984,14(6):639-643
Extracts of tissues of the lower termites, Reticulitermes flavipes and Coptotermes lacteus, and the higher termite, Nasutitermes exitiosus, possess acetyl-CoA synthetase and all the enzymes of the tricarboxylic acid cycle and are thus able to oxidize acetate to CO2. The specific activities of these enzymes in R. flavipes are sufficient to cope with the rate of acetogenesis by the gut microbiota. The presence of the malic enzyme and malate dehydrogenase, but not pyruvate carboxylase or phosphoenolpyruvate carboxykinase, indicates that they may be important as anaplerotic enzymes for the conversion of pyruvate to oxalacetate. An apparent absence of pyruvate dehydrogenase in all termites suggests that they do not convert pyruvate to acetyl-CoA, but rather convert acetate (transported from the hindgut) to this compound. All the enzymes of glycolysis were present in termite extracts. Thus any glucose absorbed from the midgut, and originating from hydrolysis of cellulose by salivary or midgut enzymes, can be metabolized by termites as an energy source.  相似文献   

11.
The anoxygenic green sulfur bacteria (GSBs) assimilate CO2 autotrophically through the reductive (reverse) tricarboxylic acid (RTCA) cycle. Some organic carbon sources, such as acetate and pyruvate, can be assimilated during the phototrophic growth of the GSBs, in the presence of CO2 or HCO3. It has not been established why the inorganic carbonis required for incorporating organic carbon for growth and how the organic carbons are assimilated. In this report, we probed carbon flux during autotrophic and mixotrophic growth of the GSB Chlorobaculum tepidum. Our data indicate the following: (a) the RTCA cycle is active during autotrophic and mixotrophic growth; (b) the flux from pyruvate to acetyl-CoA is very low and acetyl-CoA is synthesized through the RTCA cycle and acetate assimilation; (c) pyruvate is largely assimilated through the RTCA cycle; and (d) acetate can be assimilated via both of the RTCA as well as the oxidative (forward) TCA (OTCA) cycle. The OTCA cycle revealed herein may explain better cell growth during mixotrophic growth with acetate, as energy is generated through the OTCA cycle. Furthermore, the genes specific for the OTCA cycle are either absent or down-regulated during phototrophic growth, implying that the OTCA cycle is not complete, and CO2 is required for the RTCA cycle to produce metabolites in the TCA cycle. Moreover, CO2 is essential for assimilating acetate and pyruvate through the CO2-anaplerotic pathway and pyruvate synthesis from acetyl-CoA.  相似文献   

12.
The anaerobic hyperthermophilic archaea Desulfurococcus amylolyticus, Hyperthermus butylicus, Thermococcus celer, Pyrococcus woesei, the hyperthermophilic bacteria Thermotoga maritima and Clostridium thermohydrosulfuricum and the aerobic mesophilic archaeon Halobacterium saccharovorum were grown either on complex media, on sugars or on pyruvate as carbon and energy sources. During growth acetate was formed as fermentation product by all organisms. The enzymes involved in acetyl-CoA formation from pyruvate and in acetate formation from acetyl-CoA were investigated:
  1. Cell extracts of all species, both archaea and bacteria, catalyzed the coenzyme A-dependent oxidative decarboxylation of pyruvate with viologen dyes or with Clostridium pasteurianum ferredoxin as electron acceptors indicating a pyruvate: ferredoxin oxidoreductase to be operative in acetyl-CoA formation from pyruvate.
  2. Cell extracts of all archaeal species, both hyperthermophiles (D. amylolyticus, H. butylicus, T. celer, P. woesei) and the mesophile H. saccharovorum, contained an acetyl-CoA synthetase (ADP forming), which catalyzes both acetate formation from acetyl-CoA and ATP synthesis from ADP and phosphate (Pi): Acetyl-CoA+ADP+Pi?Acetate + ATP+CoA. Phosphate acetyltransferase and acetate kinase could not be detected.
  3. Cell extracts of the hyperthermophilic (eu)bacteria T. maritima and C. thermohydrosulfuricum contained phosphate acetyltransferase and acetate kinase rather than acetyl-CoA synthetase (ADP forming).
These data indicate that acetyl-CoA synthetase (ADP forming) represents a typical archaeal property rather than an enzyme specific for hyperthermophiles. It is proposed that in all acetate forming archaea the formation of acetate and of ATP from acetyl-CoA, ADP and Pi are catalyzed by acetyl-CoA synthetase (ADP forming), whereas in all acetate forming (eu)bacteria these reactions are catalyzed by two enzymes, phosphate acetyltransferase and acetate kinase.  相似文献   

13.
Mesaconate is an intermediate in the glutamate degradation pathway of microorganisms such as Clostridium tetanomorphum. However, metabolic engineering to produce mesaconate has not been reported previously. In this work, two enzymes involved in mesaconate production, glutamate mutase and 3-methylaspartate ammonia lyase from C. tetanomorphum, were recombinantly expressed in Escherichia coli. To improve mesaconate production, reactivatase of glutamate mutase was discovered and adenosylcobalamin availability was increased. In addition, glutamate mutase was engineered to improve the in vivo activity. These efforts led to efficient mesaconate production at a titer of 7.81 g/L in shake flask with glutamate feeding. Then a full biosynthetic pathway was constructed to produce mesaconate at a titer of 6.96 g/L directly from glucose. In summary, we have engineered an efficient system in E. coli for the biosynthesis of mesaconate.  相似文献   

14.
In this study, a propionate CoA-transferase (H16_A2718; EC 2.8.3.1) from Ralstonia eutropha H16 (Pct Re ) was characterized in detail. Glu342 was identified as catalytically active amino acid residue via site-directed mutagenesis. Activity of Pct Re was irreversibly lost after the treatment with NaBH4 in the presence of acetyl-CoA as it is shown for all CoA-transferases from class I, thereby confirming the formation of the covalent enzyme-CoA intermediate by Pct Re . In addition to already known CoA acceptors for Pct Re such as 3-hydroxypropionate, 3-hydroxybutyrate, acrylate, succinate, lactate, butyrate, crotonate and 4-hydroxybutyrate, it was found that glycolate, chloropropionate, acetoacetate, valerate, trans-2,3-pentenoate, isovalerate, hexanoate, octanoate and trans-2,3-octenoate formed also corresponding CoA-thioesters after incubation with acetyl-CoA and Pct Re . Isobutyrate was found to be preferentially used as CoA acceptor amongst other carboxylates tested in this study. In contrast, no products were detected with acetyl-CoA and formiate, bromopropionate, glycine, pyruvate, 2-hydroxybutyrate, malonate, fumarate, itaconate, β-alanine, γ-aminobutyrate, levulate, glutarate or adipate as potential CoA acceptor. Amongst CoA donors, butyryl-CoA, crotonyl-CoA, 3-hydroxybutyryl-CoA, isobutyryl-CoA, succinyl-CoA and valeryl-CoA apart from already known propionyl-CoA and acetyl-CoA could also donate CoA to acetate. The highest rate of the reaction was observed with 3-hydroxybutyryl-CoA (2.5 μmol mg?1 min?1). K m values for propionyl-CoA, acetyl-CoA, acetate and 3-hydroxybutyrate were 0.3, 0.6, 4.5 and 4.3 mM, respectively. The rather broad substrate range might be a good starting point for enzyme engineering approaches and for the application of Pct Re in biotechnological polyester production.  相似文献   

15.
Assimilation of acetyl coenzyme A (acetyl-CoA) is an essential process in many bacteria that proceeds via the glyoxylate cycle or the ethylmalonyl-CoA pathway. In both assimilation strategies, one of the final products is malate that is formed by the condensation of acetyl-CoA with glyoxylate. In the glyoxylate cycle this reaction is catalyzed by malate synthase, whereas in the ethylmalonyl-CoA pathway the reaction is separated into two proteins: malyl-CoA lyase, a well-known enzyme catalyzing the Claisen condensation of acetyl-CoA with glyoxylate and yielding malyl-CoA, and an unidentified malyl-CoA thioesterase that hydrolyzes malyl-CoA into malate and CoA. In this study the roles of Mcl1 and Mcl2, two malyl-CoA lyase homologs in Rhodobacter sphaeroides, were investigated by gene inactivation and biochemical studies. Mcl1 is a true (3S)-malyl-CoA lyase operating in the ethylmalonyl-CoA pathway. Notably, Mcl1 is a promiscuous enzyme and catalyzes not only the condensation of acetyl-CoA and glyoxylate but also the cleavage of β-methylmalyl-CoA into glyoxylate and propionyl-CoA during acetyl-CoA assimilation. In contrast, Mcl2 was shown to be the sought (3S)-malyl-CoA thioesterase in the ethylmalonyl-CoA pathway, which specifically hydrolyzes (3S)-malyl-CoA but does not use β-methylmalyl-CoA or catalyze a lyase or condensation reaction. The identification of Mcl2 as thioesterase extends the enzyme functions of malyl-CoA lyase homologs that have been known only as “Claisen condensation” enzymes so far. Mcl1 and Mcl2 are both related to malate synthase, an enzyme which catalyzes both a Claisen condensation and thioester hydrolysis reaction.Many organic compounds are initially metabolized to acetyl coenzyme A (acetyl-CoA), at which point they enter the central carbon metabolism. Examples of such growth substrates are C1 and C2 compounds (e.g., methanol and ethanol), fatty acids, waxes, esters, alkenes, or (poly)hydroxyalkanoates. The synthesis of all cell constituents from acetyl-CoA requires a specialized pathway for the conversion of this central C2 unit into other biosynthetic precursor metabolites. This (anaplerotic) process is referred to as acetyl-CoA assimilation, and two very different strategies have been described, i.e., the glyoxylate cycle and the ethylmalonyl-CoA pathway (12, 21) (Fig. (Fig.11).Open in a separate windowFIG. 1.Pathways for acetyl-CoA assimilation. (A) Glyoxylate cycle. The key enzymes are isocitrate lyase and malate synthase. (B) Ethylmalonyl-CoA pathway. The unique enzymes of the pathway are crotonyl-CoA carboxylase/reductase, ethylmalonyl-CoA/methylmalonyl-CoA epimerase, (2R)-ethylmalonyl-CoA mutase, (2S)-methylsuccinyl-CoA dehydrogenase, mesaconyl-CoA hydratase, (3S)-malyl-CoA/β-methylmalonyl-CoA lyase, and (3S)-malyl-CoA thioesterase. The enzymes involved in the (apparent) malate synthase reaction(s) are boxed for each pathway.The glyoxylate cycle for acetyl-CoA assimilation is in fact a modified citric acid cycle that converts two molecules of acetyl-CoA to the citric acid cycle intermediate malate (Fig. (Fig.1A)1A) (21). In a first reaction sequence, one molecule of acetyl-CoA is converted into glyoxylate due to the combined action of the initial enzymes of the citric acid cycle and isocitrate lyase, the key enzyme of this assimilation strategy. Isocitrate lyase cleaves the citric cycle intermediate isocitrate into succinate and glyoxylate (22). The glyoxylate formed is then condensed in a second step with another molecule of acetyl-CoA to yield malate and free CoA. Because the two decarboxylation reactions of the citric acid cycle are circumvented by this acetyl-CoA assimilation strategy, the glyoxylate cycle is also referred to as the “glyoxylate bypass” or “glyoxylate shunt.”The ethylmalonyl-CoA pathway for acetyl-CoA assimilation replaces the glyoxylate cycle in bacteria that lack isocitrate lyase (1, 12). In this linear pathway, three molecules of acetyl-CoA, one molecule of CO2, and one molecule of bicarbonate are converted to the citric acid cycle intermediates succinyl-CoA and malate (Fig. (Fig.1B).1B). The ethylmalonyl-CoA pathway requires at least seven unique enzymes. Crotonyl-CoA carboxylase/reductase, ethylmalonyl-CoA mutase, and methylsuccinyl-CoA dehydrogenase are considered key enzymes of the pathway, and all three enzymes have been characterized from Rhodobacter sphaeroides (12-14).Although these two acetyl-CoA strategies differ with respect to their reaction sequence, intermediates and overall balance, the glyoxylate cycle and the ethylmalonyl-CoA pathway both require the condensation of acetyl-CoA and glyoxylate to form malate (Fig. (Fig.1,1, boxed). In the glyoxylate cycle, this reaction is catalyzed by malate synthase, whereas in the ethylmalonyl-CoA pathway malate synthase is catalyzed by two separate enzymes, malyl-CoA lyase and malyl-CoA thioesterase (7, 26).Malyl-CoA lyases catalyze the reversible condensation of acetyl-CoA and glyoxylate into malyl-CoA and have been purified from Methylobacterium extorquens, Chloroflexus aurantiacus, Aminobacter aminovorans, and Rhodobacter capsulatus; the corresponding genes were identified as mclA (M. extorquens), mcl (C. aurantiacus), and mcl1 (R. capsulatus) (5, 16, 17, 19, 26). Remarkably, these proteins are promiscuous enzymes that also catalyze the (reversible) cleavage of β-methylmalyl-CoA into glyoxylate and propionyl-CoA, and it has been suggested that these enzymes catalyze both reactions in vivo (16, 19, 26). However, in contrast to malyl-CoA lyase, the malyl-CoA thioesterase catalyzing the highly exergonic hydrolysis of the CoA-thioester into malate and free CoA has not been identified so far, and the nature of the enzyme has remained enigmatic (7, 26).For R. sphaeroides, a malyl-CoA lyase homolog has been shown to be upregulated during growth on acetate, and it was proposed that this protein (Mcl1) catalyzes the cleavage of β-methylmalyl-CoA, as well as the condensation of acetyl-CoA and glyoxylate in the ethylmalonyl-CoA pathway (1). Interestingly, R. sphaeroides encodes a second malyl-CoA lyase homolog with 34% amino acid sequence identity to Mcl1. This protein, named Mcl2, was also shown to be upregulated during growth of R. sphaeroides on acetate, but a function could not be assigned so far (1). We therefore addressed the function of both malyl-CoA lyase homologs by gene inactivation and biochemical studies of recombinant Mcl1 and Mcl2. Based on our findings, we confirm here the function of Mcl1 in R. sphaeroides as (3S)-malyl-CoA/β-methylmalyl-CoA lyase and identify its paralog Mcl2 as the long-sought (3S)-malyl-CoA thioesterase.  相似文献   

16.
The mechanism of the dark assimilation of acetate in the photoheterotrophically grown nonsulfur bacterium Rhodospirillum rubrum was studied. Both in the light and in the dark, acetate assimilation in Rsp. rubrum cells, which lack the glyoxylate pathway, was accompanied by the excretion of glyoxylate into the growth medium. The assimilation of propionate was accompanied by the excretion of pyruvate. Acetate assimilation was found to be stimulated by bicarbonate, pyruvate, the C4-dicarboxylic acids of the Krebs cycle, and glyoxylate, but not by propionate. These data implied that the citramalate (CM) cycle in Rsp. rubrum cells grown aerobically in the dark can function as an anaplerotic pathway. This supposition was confirmed by respiration measurements. The respiration of cells oxidizing acetate depended on the presence of CO2 in the medium. The fact that the intermediates of the CM cycle (citramalate and mesaconate) markedly inhibited acetate assimilation but had almost no effect on cell respiration indicative that citramalate and mesaconate are intermediates of the acetate assimilation pathway. The inhibition of acetate assimilation and cell respiration by itaconate was due to its inhibitory effect on propionyl-CoA carboxylase, an enzyme of the CM cycle. The addition of 5 mM itaconate to extracts of Rsp. rubrum cells inhibited the activity of this enzyme by 85%. The data obtained suggest that the CM cycle continues to function in Rsp. rubrum cells that have been grown anaerobically in the light and then transferred to the dark and incubated aerobically.  相似文献   

17.
18.
Strain DCL14, which is able to grow on limonene as a sole source of carbon and energy, was isolated from a freshwater sediment sample. This organism was identified as a strain of Rhodococcus erythropolis by chemotaxonomic and genetic studies. R. erythropolis DCL14 also assimilated the terpenes limonene-1,2-epoxide, limonene-1,2-diol, carveol, carvone, and (−)-menthol, while perillyl alcohol was not utilized as a carbon and energy source. Induction tests with cells grown on limonene revealed that the oxygen consumption rates with limonene-1,2-epoxide, limonene-1,2-diol, 1-hydroxy-2-oxolimonene, and carveol were high. Limonene-induced cells of R. erythropolis DCL14 contained the following four novel enzymatic activities involved in the limonene degradation pathway of this microorganism: a flavin adenine dinucleotide- and NADH-dependent limonene 1,2-monooxygenase activity, a cofactor-independent limonene-1,2-epoxide hydrolase activity, a dichlorophenolindophenol-dependent limonene-1,2-diol dehydrogenase activity, and an NADPH-dependent 1-hydroxy-2-oxolimonene 1,2-monooxygenase activity. Product accumulation studies showed that (1S,2S,4R)-limonene-1,2-diol, (1S,4R)-1-hydroxy-2-oxolimonene, and (3R)-3-isopropenyl-6-oxoheptanoate were intermediates in the (4R)-limonene degradation pathway. The opposite enantiomers [(1R,2R,4S)-limonene-1,2-diol, (1R,4S)-1-hydroxy-2-oxolimonene, and (3S)-3-isopropenyl-6-oxoheptanoate] were found in the (4S)-limonene degradation pathway, while accumulation of (1R,2S,4S)-limonene-1,2-diol from (4S)-limonene was also observed. These results show that R. erythropolis DCL14 metabolizes both enantiomers of limonene via a novel degradation pathway that starts with epoxidation at the 1,2 double bond forming limonene-1,2-epoxide. This epoxide is subsequently converted to limonene-1,2-diol, 1-hydroxy-2-oxolimonene, and 7-hydroxy-4-isopropenyl-7-methyl-2-oxo-oxepanone. This lactone spontaneously rearranges to form 3-isopropenyl-6-oxoheptanoate. In the presence of coenzyme A and ATP this acid is converted further, and this finding, together with the high levels of isocitrate lyase activity in extracts of limonene-grown cells, suggests that further degradation takes place via the β-oxidation pathway.  相似文献   

19.
Nicotine is an important environmental toxicant in tobacco waste. Shinella sp. strain HZN7 can metabolize nicotine into nontoxic compounds via variations of the pyridine and pyrrolidine pathways. However, the catabolic mechanism of this variant pathway at the gene or enzyme level is still unknown. In this study, two 6-hydroxynicotine degradation-deficient mutants, N7-M9 and N7-W3, were generated by transposon mutagenesis. The corresponding mutant genes, designated nctB and tnp2, were cloned and analyzed. The nctB gene encodes a novel flavin adenine dinucleotide-containing (S)-6-hydroxynicotine oxidase that converts (S)-6-hydroxynicotine into 6-hydroxy-N-methylmyosmine and then spontaneously hydrolyzes into 6-hydroxypseudooxynicotine. The deletion and complementation of the nctB gene showed that this enzyme is essential for nicotine or (S)-6-hydroxynicotine degradation. Purified NctB could also convert (S)-nicotine into N-methylmyosmine, which spontaneously hydrolyzed into pseudooxynicotine. The kinetic constants of NctB toward (S)-6-hydroxynicotine (Km = 0.019 mM, kcat = 7.3 s−1) and nicotine (Km = 2.03 mM, kcat = 0.396 s−1) indicated that (S)-6-hydroxynicotine is the preferred substrate in vivo. NctB showed no activities toward the R enantiomer of nicotine or 6-hydroxynicotine. Strain HZN7 could degrade (R)-nicotine into (R)-6-hydroxynicotine without any further degradation. The tnp2 gene from mutant N7-W3 encodes a putative transposase, and its deletion did not abolish the nicotine degradation activity. This study advances the understanding of the microbial diversity of nicotine biodegradation.  相似文献   

20.
Brightman, Vernon (The University of Chicago, Chicago), and William R. Martin. Pathway for the dissimilation of itaconic and mesaconic acids. J. Bacteriol. 82:376-382. 1961.-Studies on the oxidation of itaconic and mesaconic acids by a Pseudomonas sp., adapted to utilize either of these acids as a sole carbon source, have provided evidence for a pathway converting both itaconate and mesaconate to succinate. A metabolic interconversion of itaconate, mesaconate, and citramalate has also been demonstrated by whole cell and cell-free enzyme studies.Succinate derived from methylene-labeled itaconate was found to be labeled in the inside carbon atoms, a fact which indicates that the branched chain compound was converted into a straight chain molecule by a shift of the methylene carbon (C-5) from the side chain of itaconate to a position between C-2 and C-3 in an, as yet, unknown straight chain intermediate prior to its conversion to succinate.  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