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.
In attempting to determine the nature of the enzyme system mediating the conversion of catechol to diphenylenedioxide 2,3-quinone, in Tecoma leaves, further purification of the enzyme was undertaken. The crude enzyme from Tecoma leaves was processed further by protamine sulfate precipitation, positive adsorption on tricalcium phosphate gel, and elution and chromatography on DEAE-Sephadex. This procedure yielded a 120-fold purified enzyme which stoichiometrically converted catechol to diphenylenedioxide 2,3-quinone. The purity of the enzyme system was assessed by polyacrylamide gel electrophoresis. The approximate molecular weight of the enzyme was assessed as 200,000 by gel filtration on Sephadex G-150. The enzyme functioned optimally at pH 7.1 and at 35 °C. The Km for catechol was determined as 4 × 10?4m. The enzyme did not oxidize o-dihydric phenols other than catechol and it did not exhibit any activity toward monohydric and trihydric phenols and flavonoids. Copper-chelating agents did not inhibit the enzyme activity. Copper could not be detected in the purified enzyme preparations. The purified enzyme was not affected by extensive dialysis against copper-complexing agents. It did not show any peroxidase activity and it was not inhibited by catalase. Hydrogen peroxide formation could not be detected during the catalytic reaction. The enzymatic conversion of catechol to diphenylenedioxide 2,3-quinone by the purified Tecoma leaf enzyme was suppressed by such reducing agents as GSH and cysteamine. The purified enzyme was not sensitive to carbon monoxide. It was not inhibited by thiol inhibitors. The Tecoma leaf was found to be localized in the soluble fraction of the cell. Treatment of the purified enzyme with acid, alkali, and urea led to the progressive denaturation of the enzyme.  相似文献   

2.
Fermentative redox balance has long been utilized as a metabolic evolution platform to improve efficiency of NADH-dependent pathways. However, such system relies on the complete recycling of NADH and may become limited when the target pathway results in excess NADH stoichiometrically. In this study, endogenous capability of Escherichia coli for 2,3-butanediol (2,3-BD) synthesis was explored using the anaerobic selection platform based on redox balance. To address the issue of NADH excess associated with the 2,3-BD pathway, we devised a substrate-decoupled system where a pathway intermediate is externally supplied in addition to the carbon source to decouple NADH recycling ratio from the intrinsic pathway stoichiometry. In this case, feeding of the 2,3-BD precursor acetoin effectively restored anaerobic growth of the mixed-acid fermentation mutant that remained otherwise inhibited even in the presence of a functional 2,3-BD pathway. Using established 2,3-BD dehydrogenases as model enzyme, we verified that the redox-based selection system is responsive to NADPH-dependent reactions but with lower sensitivity. Based on this substrate-decoupled selection scheme, we successfully identified the glycerol/1,2-propanediol dehydrogenase (Ec-GldA) as the major enzyme responsible for the acetoin reducing activity (kcat/Km≈0.4 mM−1 s−1) observed in E. coli. Significant shift of 2,3-BD configuration upon withdrawal of the heterologous acetolactate decarboxylase revealed that the endogenous synthesis of acetoin occurs via diacetyl. Among the predicted diacetyl reductase in E. coli, Ec-UcpA displayed the most significant activity towards diacetyl reduction into acetoin (Vmax≈6 U/mg). The final strain demonstrated a meso-2,3-BD production titer of 3 g/L without introduction of foreign genes. The substrate-decoupled selection system allows redox balance regardless of the pathway stoichiometry thus enables segmented optimization of different reductive pathways through enzyme bioprospecting and metabolic evolution.  相似文献   

3.
Indoleamine 2,3-dioxygenase (IDO) and tryptophan 2,3-dioxygenase (TDO) are tryptophan-degrading enzymes that catalyze the same reaction, the first step in tryptophan catabolism via the kynurenine pathway. TDO is widely distributed among life-forms, being found not only in eukaryotes but also in bacteria. In contrast, IDO has been found only in mammals and yeast to date. However, recent genome and EST projects have identified IDO homologues in non-mammals and found an IDO paralogue that is expressed in mice. In this study, we cloned the frog and fish IDO homologues and the mouse IDO paralogue, and characterized their enzymatic properties using recombinants. The IDOs of lower vertebrates and the mouse IDO paralogue had IDO activity but had 500–1000 times higher K m values and very low enzyme efficiency compared with mammalian IDOs. It appears that L-Trp is not a true substrate for these enzymes in vivo, although their actual function is unknown. On the phylogenetic tree, these low-activity IDOs, which we have named “proto-IDOs,” formed a cluster that was distinct from the mammalian IDO cluster. The IDO and proto-IDO genes are present tandemly on the chromosomes of mammals, including the marsupial opossum, whereas only the proto-IDO gene is observed in chicken and fish genomes. These results suggest that (mammalian) IDOs arose from proto-IDOs by gene duplication that occurred before the divergence of marsupial and eutherian (placental) mammals in mammalian evolutionary history.  相似文献   

4.
The activity and expression of indoleamine 2,3-dioxygenase together with l-tryptophan transport has been examined in cultured human breast cancer cells. MDA-MB-231 but not MCF-7 cells expressed mRNA for indoleamine 2,3-dioxygenase. Kynurenine production by MDA-MB-231 cells, which was taken as a measure of enzyme activity, was markedly stimulated by interferon-γ (1000 units/ml). Accordingly, l-tryptophan utilization by MDA-MB-231 cells was enhanced by interferon-γ. 1-Methyl-dl-tryptophan (1 mM) inhibited interferon-γ induced kynurenine production by MBA-MB-231 cells. Kynurenine production by MCF-7 cells remained at basal levels when cultured in the presence of interferon-γ. l-Tryptophan transport into MDA-MB-231 cells was via a Na+-independent, BCH-sensitive pathway. It appears that system L (LAT1/CD98) may be the only pathway for l-tryptophan transport into these cells. 1-Methyl-d,l-tryptophan trans-stimulated l-tryptophan efflux from MDA-MB-231 cells and thus appears to be a transported substrate of system L. The results suggest that system L plays an important role in providing indoleamine-2,3-dioxygenase with its main substrate, l-tryptophan, and suggest a mechanism by which estrogen receptor-negative breast cancer cells may evade the attention of the immune system.  相似文献   

5.
Furo[2,3-f]quinolin-5-ol derivatives were synthesized and tested for in vitro antifungal activity against Candida,Aspergillus species, and Cryptococcus neoformans. Among them tested, many furo[2,3-f]quinolin-5-ols showed good antifungal activity. The results suggest that furo[2,3-f]quinolin-5-ols would be promising antifungal agents.  相似文献   

6.
In our previous paper (Biochim. Biophys. Acta 1379 (1998) 257–263), we demonstrated that bicarbonate promotes a cleavage of lactone ring of dehydroascorbate (DHA) on the basis of in vitro experiments. In the present study, we examined the degradation of DHA in blood circulation in vivo by using a high-performance liquid chromatographic method for the determination of ascorbate (AsA), DHA and 2,3-diketogulonate (2,3-DKG), which required no pretreatment of biological fluids. When DHA was intravenously administered to rats, a rapid disappearance of DHA (t1/2<1 min) and a concomitant appearance of 2,3-DKG in blood circulation were observed. Approximately 90% of the administered DHA were excreted into urine as resulting 2,3-DKG (55%) and AsA (31%), respectively. Furthermore, we elucidated that rat plasma lacks an enzyme having an aldonolactonase-like activity. The result of the present study suggests that this DHA disappearance is a function of both a chemical degradation to 2,3-DKG and a reduction to AsA.  相似文献   

7.
Using a conserved sequence motif, a new gene (YAL060W) of the MDR family has been identified in Saccharomyces cerevisiae. The expressed protein was a stereoespecific (2R,3R)-2,3-butanediol dehydrogenase (BDH). The best substrates were (2R,3R)-2,3-butanediol for the oxidation and (3R/3S)-acetoin and 1-hydroxy-2-propanone for the reduction reactions. The enzyme is extremely specific for NAD(H) as cofactor, probably because the presence of Glu223 in the cofactor binding site, instead of the highly conserved Asp223. BDH is inhibited competitively by 4-methylpyrazole with a Ki of 34 μM. Yeast could grow on 2,3-butanediol or acetoin as a sole energy and carbon sources, and a 3.6-fold increase in BDH activity was observed when cells were grown in 2,3-butanediol, suggesting a role of the enzyme in 2,3-butanediol metabolism. However, the disruption of the YAL060W gene was not lethal for the yeast under laboratory conditions, and the disrupted strain could also grow in 2,3-butanediol and acetoin. This suggests that other enzymes, in addition to BDH, can also metabolize 2,3-butanediol in yeast.  相似文献   

8.
Bisphosphoglycerate synthase from horse red cells has been purified to apparent homogeneity by a simple and efficient new procedure incorporating chromatography on a column of Sepharose 4B derivatized with blue dextran. The enzyme is similar to the human red cell synthase in subunit size. It is phosphorylated by either glycerate-1,3-P2 or glycerate-2,3-P2 to form a phosphoenzyme with the acid-lability of a histidyl phosphate. In addition to the synthase activity (glycerate-1,3-P2 → glycerate-2,3-P2), kcat 12.5 s?1, the enzyme has bisphosphoglycerate phosphatase activity in the presence of glycolate-2-P (glycerate-2,3-P2 → glycerate-P + Pi), kcat 2.6 s?1 and phosphoglycerate mutase activity (3-PGA ? 2-PGA), kcat 1.7 s?1. The energy of activation for the synthase reaction is 9.38 kcal/mol. Lineweaver-Burk plots of the kinetic data are parallel lines. In contrast intersecting patterns were obtained from similar experiments done with the human red cell enzyme. Further investigation is required to explain these differences. This enzyme may function as both synthase and phosphatase for bisphosphoglycerate in the red blood cell.  相似文献   

9.
2,3-Diphosphoglycerate (2,3-DPG), an intracellular metabolite of glycolytic pathway is known to affect the oxygen binding capacity of haemoglobin and mechanical properties of the red blood cells. 2,3-DPG levels have been reported to be elevated during anaemic conditions including visceral leishmaniasis. 2,3-DPG activity in P. falciparum infected red blood cells, particularly in cells infected with different stages of the parasite and its relationship with structural integrity of the cells is not known. Chloroquine sensitive and resistant strains of P. falciparum were cultured in vitro and synchronized cultures of ring, trophozoite and schizont stage rich cells along with the uninfected control erythrocytes were assayed for 2,3-DPG activity and osmotic fragility. It was observed that in both the strains, in infected erythrocytes the 2,3-DPG activity gradually decreased and osmotic fragility gradually increased as the parasite matured from ring to schizont stage. The decrease in 2,3-DPG may probably be due to increased pyruvate kinase activity of parasite origin, which has been shown in erythrocytes infected with several species of Plasmodium. The absence of compensatory increase in 2,3-DPG in P. falciparum infected erythrocytes may aggravate hypoxia due to anaemia in malaria and probably may contribute to hypoxia in cerebral malaria. As 2,3-DPG was not found to be increased in erythrocytes parasitized with P. falciparum, the increased osmotic fragility observed in these cells is not due to increased 2,3-DPG as has been suggested in visceral leishmaniasis.  相似文献   

10.
In addition to hyoscine and hyoscyamine, a new compound isolated from the leaves and stems of Hyoscyamus albus has been characterized as 2,3-dimethylnonacosane by spectral studies.  相似文献   

11.
The initial concentration of corn steep liquor (CSL) have remarkable effects on not only 2,3-butanediol (2,3-BD) and acetoin (metabolic precursor) production, but also on the ratio of 2,3-BD to acetoin. When a high concentration of CSL was supplemented, cell growth was improved, acetoin reductase (ACR) was stimulated, the concentration of 2,3-BD increased by 78.6%, acetoin decreased by 61.9%, and the ratio of 2,3-BD to acetoin increased by 3.69-fold. The acr gene, encoding ACR, was over-expressed in Bacillus subtilis. Compared to the control (parent strain), low levels of CSL in the engineered strain increased 2,3-BD concentration and the ratio 2,3-BD to acetoin by 13.9% and 39.5%, respectively, and decreased acetoin titer by 18.3%. Acetoin became a major product under low levels of CSL. Also, a knockout strain carrying an acr::cat insertion mutation was constructed. As expected, the loss of ACR activity led to an accumulation of acetoin in the supernatants of acr:: cat mutant cultures. Additionally, the productivity of acetoin was improved by high concentration of CSL. The results above demonstrate the feasibility of using B. subtilis for the production of not only 2,3-BD but also acetoin as a major product.  相似文献   

12.
The production of biofuels by recombinant Escherichia coli is restricted by the toxicity of the products. 2,3-Butanediol (2,3-BD), a platform and fuel bio-chemical with low toxicity to microbes, could be a promising alternative for biofuel production. However, the yield and productivity of 2,3-BD produced by recombinant E. coli strains are not sufficient for industrial scale fermentation. In this work, the production of 2,3-BD by recombinant E. coli strains was optimized by applying a systematic approach. 2,3-BD biosynthesis gene clusters were cloned from several native 2,3-BD producers, including Bacillus subtilis, Bacillus licheniformis, Klebsiella pneumoniae, Serratia marcescens, and Enterobacter cloacae, inserted into the expression vector pET28a, and compared for 2,3-BD synthesis. The recombinant strain E. coli BL21/pETPT7-EcABC, carrying the 2,3-BD pathway gene cluster from Enterobacter cloacae, showed the best ability to synthesize 2,3-BD. Thereafter, expression of the most efficient gene cluster was optimized by using different promoters, including PT7, Ptac, Pc, and Pabc. E. coli BL21/pET-RABC with Pabc as promoter was superior in 2,3-BD synthesis. On the basis of the results of biomass and extracellular metabolite profiling analyses, fermentation conditions, including pH, agitation speed, and aeration rate, were optimized for the efficient production of 2,3-BD. After fed-batch fermentation under the optimized conditions, 73.8 g/L of 2,3-BD was produced by using E. coli BL21/pET-RABC within 62 h. The values of both yield and productivity of 2,3-BD obtained with the optimized biological system are the highest ever achieved with an engineered E. coli strain. In addition to the 2,3-BD production, the systematic approach might also be used in the production of other important chemicals through recombinant E. coli strains.  相似文献   

13.
Pseudomonas putida strain BNF1 was isolated to degrade aromatic hydrocarbons efficiently and use phenol as a main carbon and energy source to support its growth. Catechol 2,3-dioxygenase was found to be the responsible key enzyme for the biodegradation of aromatic hydrocarbons. Catechol 2,3-dioxygenase gene was cloned from plasmid DNA of P. putida strain BNF1. The nucleotide base sequence of a 924 bp segment encoding the catechol 2,3-dioxygenase (C23O) was determined. This segment showed an open reading frame, which encoded a polypeptide of 307 amino acids. C23O gene was inserted into NotI-cut transposon vector pUT/mini-Tn5 (Kmr) to get a novel transposon vector pUT/mini-Tn5-C23O. With the helper plasmid PRK2013, the transposon vector pUT/mini-Tn5-C23O was introduced into one alkanes degrading strain Acinetobacter sp. BS3 by triparental conjugation, and then the C23O gene was integrated into the chromosome of Acinetobacter sp. BS3. And the recombinant BS3-C23O, which could express catechol 2,3-dioxygenase protein, was obtained. The recombinant BS3-C23O was able to degrade various aromatic hydrocarbons and n-alkanes. Broad substrate specificity, high enzyme activity, and the favorable stability suggest that the BS3-C23O was a potential candidate used for the biodegradation of crude oil.  相似文献   

14.
Yuasa HJ  Ushigoe A  Ball HJ 《Gene》2011,485(1):22-31
Indoleamine 2,3-dioxygenase (IDO) and tryptophan 2,3-dioxygenase (TDO) are tryptophan-degrading enzymes that catalyze the first step in L-Trp catabolism via the kynurenine pathway. In mammals, TDO is mainly expressed in the liver and primarily supplies nicotinamide adenine dinucleotide (NAD+). TDO is widely distributed from mammals to bacteria. Active IDO enzymes have been reported only in vertebrates and fungi. In mammals, IDO activity plays a significant role in the immune system while in fungal species, IDO is constitutively expressed and supplies NAD+, like mammalian TDO. A search of genomic databases reveals that some bacterial species also have a putative IDO gene. A phylogenetic analysis clustered bacterial IDOs into two groups, group I or group II bacterial IDOs. The catalytic efficiencies of group I bacterial IDOs were very low and they are suspected not to contribute significantly to L-Trp metabolism. The bacterial species bearing the group I bacterial IDO are scattered across a few phyla and no phylogenetically close relationship is observed between them. This suggests that the group I bacterial IDOs might be acquired by horizontal gene transmission that occurred in each lineage independently. In contrast, group II bacterial IDOs showed rather high catalytic efficiency. Particularly, the enzymatic characteristics (Km, Vmax and inhibitor selectivity) of the Gemmatimonas aurantiaca IDO are comparable to those of mammalian IDO1, although comparison of the IDO sequences does not suggest a close evolutionary relationship. In several bacteria, TDO and the kynureninase gene (kynU) are clustered on their chromosome suggesting that these genes could be transcribed in an operon. Interestingly, G. aurantiaca has no TDO, and the IDO is clustered with kynU on its chromosome. Although the G. aurantiaca also has NadA and NadB to synthesize a quinolinic acid (a precursor of NAD+) via the aspartate pathway, the high activity of the G. aurantiaca IDO flanking the kynU gene suggests its IDO has a function similar to eukaryotic enzymes.  相似文献   

15.
Bacillus subtilis is capable of producing 2,3-butanediol from acetoin by fermentation, but to date, the gene encoding the enzyme responsible, acetoin reductase/2,3-butanediol dehydrogenase (AR/BDH), has remained unknown. A search of the B. subtilis genome database with the amino acid sequences of functional AR/BDHs from Saccharomyces cerevisiae and Bacillus cereus resulted in the identification of a highly similar protein encoded by the B. subtilis ydjL gene. A knockout strain carrying a ydjL::cat insertion mutation was constructed, which (i) abolished 2,3-butanediol production in early stationary phase, (ii) produced no detectable AR or BDH activity in vitro, and (iii) accumulated the precursor acetoin in early stationary phase. The ydjL::cat mutation also affected the kinetics of lactate but not acetate production during stationary-phase cultivation with glucose under oxygen limitation. A very small amount of 2,3-butanediol was detected in very-late-stationary-phase (96-hour) cultures of the ydjL::cat mutant, suggesting the existence of a second gene encoding a minor AR activity. From the data, it is proposed that the major AR/BDH-encoding gene ydjL be renamed bdhA.  相似文献   

16.
《Insect Biochemistry》1989,19(5):457-461
The properties of xanthine dehydrogenase from Drosophila virilis have been studied and the effects of adenine on the enzyme analysed. The enzyme has a much higher thermal stability than that from D. melanogaster. Ammonium sulphate fractionation of the extracts precipitated the enzyme in the 0–40% saturation stage whereas the D. melanogaster enzyme precipitates between 40–60% saturation. Xanthine dehydrogenase activity per mg protein in crude extracts is 40–45% lower in D. virilis than in D. melanogaster but it is more resistant to dialysis. In vitro addition of adenine inhibits both enzymes but does not appear to stabilise the D. virilis enzyme during dialysis.  相似文献   

17.
The levels of cyclic 2,3-diphosphoglycerate (cDPG) in methanogenic bacteria are governed by the antagonistic activities of cDPG synthetase and cDPG hydrolase. In this paper we focus on the synthetase from Methanobacterium thermoautotrophicum. The cytoplasmic 150 kDa enzyme catalyzed cDPG synthesis from 2,3-diphosphoglycerate (apparent Km=21 mM), Mg2+ (Km=3.1 mM) and ATP (Km=1–2 mM). In batch-fed cultures, the enzyme was constitutively present (6–6.5 nmol per min per mg protein) during the different growth phases. In continuous cultures, activity decreased in response to phosphate limitation. The synthetase reaction proceeded with maximal rate at pH 6 and at 65° C and was specifically dependent on high (>0.3M) K+ concentrations. The reaction conditions remarkably contrasted to those of cDPG degradation catalyzed by the previously described membrane-bound cDPG hydrolase.Abbreviations cDPG Cyclic 2,3-diphosphoglycerate - 2,3-DPG 2,3-Diphosphoglycerate - 2-PG 2-Phosphoglycerate - 3-PG 3-Phosphoglycerate  相似文献   

18.
Squalene-2,3-epoxide-cycloartenol cyclase and cycloeucalenol-obtusifoliol isomerase activities were found in microsomal fractions of corn (Zea mays) embryos. Squalene-2,3-epoxide, 1-trans-1′-norsqualene-2,3-epoxide and 1-cis-1′-norsqualene-2,3-epoxide were incubated. Squalene-2,3-epoxide was cyclized giving only cycloartenol, whereas 1-trans-1′-norsqualene-2,3-epoxide gave 31-norcycloartenol and 31-norlanosterol with a reduced yield, 1-cis-1′-norsqualene-2,3-epoxide was not significantly cyclized.  相似文献   

19.
Acetoin is widely used in food and cosmetics industries as a taste and fragrance enhancer. To produce (R)-acetoin in Saccharomyces cerevisiae, acetoin biosynthetic genes encoding α-acetolactate synthase (AlsS) and α-acetolactate decarboxylase (AlsD) from Bacillus subtilis and water-forming NADH oxidase (NoxE) from Lactococcus lactis were integrated into delta-sequences in JHY605 strain, where the production of ethanol, glycerol, and (R,R)-2,3-butanediol (BDO) was largely eliminated. We further improved acetoin production by increasing acetoin tolerance by adaptive laboratory evolution, and eliminating other byproducts including meso-2,3-BDO and 2,3-dimethylglycerate, a newly identified byproduct. Ara1, Ypr1, and Ymr226c (named Ora1) were identified as (S)-alcohol-forming reductases, which can reduce (R)-acetoin to meso-2,3-BDO in vitro. However, only Ara1 and Ypr1 contributed to meso-2,3-BDO production in vivo. We elucidate that Ora1, having a substrate preference for (S)-acetoin, reduces (S)-α-acetolactate to 2,3-dimethylglycerate, thus competing with AlsD-mediated (R)-acetoin production. By deleting ARA1, YPR1, and ORA1, 101.3 g/L of (R)-acetoin was produced with a high yield (96% of the maximum theoretical yield) and high stereospecificity (98.2%).  相似文献   

20.
Indoleamine 2,3-dioxygenase activity in the supernatant fractions (30,000g, 30 min) from various tissues of mice increased almost linearly after a single intraperitoneal administration of bacterial lipopolysaccharide (5 to 20 μg/mouse). The most prominent effect was observed in the lung, where both specific and total enzyme activities increased 40 to 80-fold during the first 24 h. Significant (10- to 20-fold) stimulation was also observed in the seminal vesicle, coagulating gland, colon, and caecum, and severalfold in the trachea, stomach, heart, small intestine, and spleen. Lipid A fraction, the biologically active unit in the lipopolysaccharide complex, was as active as the lipopolysaccharide preparations from either Escherichia coli or Salmonella S and R mutant strains, whereas the polysaccharide fraction was inactive under identical experimental conditions. When mice were pretreated with a series of daily injections of bacterial lipopolysaccharide, enzyme induction was no longer evident, indicating that tolerance to this agent had developed and that enzyme induction was caused by lipopolysaccharide but not by possible contaminants in the preparations. The enzyme activities from normal and lipopolysaccharide-treated mice were exclusively found in the soluble fractions of mouse lung homogenates. Other enzyme activities in the lung such as lysosomal (acid phosphatase), microsomal (prostaglandin cyclooxygenase), mitochondrial (monoamine oxidase and superoxide dismutase), and soluble enzyme activities (lipooxygenase and superoxide dismutase) were not significantly altered by this treatment. This increase in the enzyme activity with the lipopolysaccharide treatment was abolished with a simultaneous administration of cycloheximide or actinomycin D, and an immunological analysis with antibody for mouse enzyme (rabbit IgG) demonstrated that the observed increment of the enzyme activity was essentially due to an increase in the enzyme protein.  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