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 46 毫秒
1.
Degradation of indole by an indole-degrading methanogenic consortium enriched from sewage sludge proceeded through a two-step hydroxylation pathway yielding oxindole and isatin. The ability of this consortium to hydroxylate and subsequently degrade substituted indoles was investigated. Of the substituted indoles tested, the consortium was able to transform or degrade 3-methylindole and 3-indolyl acetate. Oxindole, 3-methyloxindole, and indoxyl were identified as metabolites of indole, 3-methylindole, and 3-indolyl acetate degradation, respectively. Isatin (indole-2,3-dione) was produced as an intermediate when the consortium was amended with oxindole, providing evidence that degradation of indole proceeded through successive hydroxylation of the 2- and 3-positions prior to ring cleavage between the C-2 and C-3 atoms on the pyrrole ring of indole. The presence of a methyl group (-CH3) at either the 1- or 2-position of indole inhibited the initial hydroxylation reaction. The substituted indole, 3-methylindole, was hydroxylated in the 2-position but not in the 3-position and could not be further metabolized through the oxindole-isatin pathway. Indoxyl (indole-3-one), the deacetylated product of 3-indolyl acetate, was not hydroxylated in the 2-position and thus was not further metabolized by the consortium. When an H atom or electron-donating group (i.e., -CH3) was present at the 3-position, hydroxylation proceeded at the 2-position, but the presence of electron-withdrawing substituent groups (i.e., -OH or -COOH) at the 3-position inhibited hydroxylation.  相似文献   

2.
Degradation of indole by an indole-degrading methanogenic consortium enriched from sewage sludge proceeded through a two-step hydroxylation pathway yielding oxindole and isatin. The ability of this consortium to hydroxylate and subsequently degrade substituted indoles was investigated. Of the substituted indoles tested, the consortium was able to transform or degrade 3-methylindole and 3-indolyl acetate. Oxindole, 3-methyloxindole, and indoxyl were identified as metabolites of indole, 3-methylindole, and 3-indolyl acetate degradation, respectively. Isatin (indole-2,3-dione) was produced as an intermediate when the consortium was amended with oxindole, providing evidence that degradation of indole proceeded through successive hydroxylation of the 2- and 3-positions prior to ring cleavage between the C-2 and C-3 atoms on the pyrrole ring of indole. The presence of a methyl group (-CH3) at either the 1- or 2-position of indole inhibited the initial hydroxylation reaction. The substituted indole, 3-methylindole, was hydroxylated in the 2-position but not in the 3-position and could not be further metabolized through the oxindole-isatin pathway. Indoxyl (indole-3-one), the deacetylated product of 3-indolyl acetate, was not hydroxylated in the 2-position and thus was not further metabolized by the consortium. When an H atom or electron-donating group (i.e., -CH3) was present at the 3-position, hydroxylation proceeded at the 2-position, but the presence of electron-withdrawing substituent groups (i.e., -OH or -COOH) at the 3-position inhibited hydroxylation.  相似文献   

3.
After enrichment culture with indole-3-carboxylate in static culture, a novel reversible decarboxylase, indole-3-carboxylate decarboxylase, was found in Arthrobacter nicotianae FI1612 and several molds. The enzyme reaction was examined in resting-cell reactions with A. nicotianae FI1612. The enzyme activity was induced specifically by indole-3-carboxylate, but not by indole. The indole-3-carboxylate decarboxylase of A. nicotianae FI1612 catalyzed the nonoxidative decarboxylation of indole-3-carboxylate into indole, and efficiently carboxylated indole and 2-methylindole by the reverse reaction. In the presence of 1 mM dithiothreitol, 50 mM Na2 S2O3, and 20% (v/v) glycerol, indole-3-carboxylate decarboxylase was partially purified from A. nicotianae FI1612. The purified enzyme had a molecular mass of approximately 258 kDa. The enzyme did not need any cofactor for the decarboxylating and carboxylating reactions.  相似文献   

4.
The tryptophan synthase alpha 2 beta 2 complex catalyzes tryptophan (Trp) biosynthesis from serine plus either indole (IN) or indole-3-glycerol phosphate (InGP). The photoreactive 5-azido analog in IN (AzIN), itself a substrate in the dark, was utilized to examine the substrate binding sites on this enzyme. When irradiated with AzIN at concentrations approaching IN saturation for the IN----Trp activity (0.1 mM), in the absence of serine, the enzyme was increasingly inactivated (up to 70-80%) concomitant with the progressive binding of a net of 2 mol AzIN per alpha beta equivalent. Little or no cooperativity in the binding of the 2 mol AzIN was observed. In contrast, there was minimal effect on the IN----InGP activity. Under these conditions AzIN appeared to be incorporated equally into each subunit. No significant inactivation nor binding occurred in the presence of serine. A quantitatively similar inactivation of InGP----Trp activity was observed over the same AzIN concentration range, suggesting common IN sites for Trp biosynthesis from either indole substrate. At higher concentrations (0.1-0.7 mM), no further inactivation occurred, although there was extensive additional binding (up to 10 mol/alpha beta equivalent). These data are consistent, although more clear-cut quantitatively, with the high- and low-affinity sites proposed from equilibrium dialysis studies. AzIN binding studies utilizing the isolated beta 2 subunit confirmed earlier reports suggesting the existence of many nonspecific IN binding sites on this subunit.  相似文献   

5.
The reaction of tryptamine with indolyl-3-alkane alpha-hydroxylase is shown to remove stereospecifically the pro-S hydrogen at C-2 of the side chain and to give hydroxytryptamine of "R" configuration. The reaction therefore proceeds stereospecifically with net inversion of configuration at C-2 of the tryptamine side chain. In the reaction of L-tryptophan methyl ester, the enzyme also catalyzes stereospecific removal of the pro-S hydrogen at C-3, but the product 3-hydroxytryptophan methyl ester is racemic at C-3. The unreacted tryptophan methyl ester is shown to incorporate solvent hydrogen into the pro-S position at C-3 in an at least partially stereospecific manner, suggesting that the reaction of L-tryptophan methyl ester is reversible. The hydrogens at C-1 of the tryptamine side chain and the alpha-hydrogen of L-tryptophan methyl ester are shown to be retained in the reactions. The results support the notion that the enzyme catalyzes stereospecific 1,4-dehydrogenation of 3-substituted indoles to the coresponding alkylidene indolenines as the primary reaction, followed by stereospecific or nonstereospecific hydration of these intermediates as a secondary process. Substrate specificity studies with a number of tryptophan analogs are in excellent agreement with such a mechanism.  相似文献   

6.
Indole reacts with sodium nitrite and glycine-HCl buffer, pH 2.6, to form a red color that is stable for more than 1 week. The reaction is reproducible and is linear over a wide range of indole concentrations (0.05–1.00 μmol). Twelve indole derivatives, including tryptophan, and 17 protein amine acids do not interfere. Indole-3-acetic acid, indole-3-acrylic acid, indole-3-pyruvic acid, 5-indole carboxylic acid, and 5-hydroxyindole-3-acetic acid interfere to varying extents (16–27%). Free indole was determined in biological material containing tryptophan by the present method. The method is also applicable to the assay of tryptophanase activity without prior indole extraction.  相似文献   

7.
Phorbol 12-myristate 13-acetate (100 nM), a potent protein kinase C and macrophage activator, has a biphasic affect on 25(OH)D3-1 alpha-hydroxylase activity in synovial fluid macrophages from arthritis patients. After 5 h, 1 alpha, 25(OH)D3 synthesis fell from 5.2 +/- 0.1 to 1.6 +/- 0.2 pmol/h per 10(6) cells, however, after 24 h and 48 h, synthesis increased to 17.4 +/- 0.3 and 22.3 +/- 1.4 pmol/h per 10(6) cells, respectively. Although an independent short-term mechanism is suggested, protein kinase C may promote macrophage activation, thus increasing long-term 25(OH)D3-1 alpha-hydroxylase expression. Intracellular calcium and cAMP are unlikely to activate the enzyme, since 0.1 microM of the calcium ionophore, A23187, and 1 mM dibutyryl-cAMP inhibited synthesis by 87% and 79%, respectively, after 24 h.  相似文献   

8.
Some strains of Bradyrhizobium japonicum have the ability to catabolize indole-3-acetic acid (IAA). Examination of this catabolism in strain 110 by in vivo experiments has revealed an enzymatic activity catalyzing the degradation of IAA and 5-hydroxy-indole-3-acetic acid. The activity requires addition of the substrates for induction and is oxygen dependent. The highest activity is obtained when the concentration of inducer is 0.2 mM. Spectrophotometric data are consistent with the suggestion that the indole ring is broken during degradation of IAA. We hypothesize that the enzyme catalyzes an oxygen-consuming opening of the indole ring analogous to the one catalyzed by tryptophan 2,3-dioxygenase. The pattern of metabolite usage by known tryptophan-auxotrophic mutants and studies of metabolites by high-performance liquid chromatography indicate that anthranilic acid is a terminal degradation product in the proposed pathway.  相似文献   

9.
Trifluoroalanine is a mechanism-based inactivator of Escherichia coli tryptophan indole-lyase (tryptophanase) and E. coli tryptophan synthase (R. B. Silverman and R. H. Abeles, 1976, Biochemistry 15, 4718-4723). We have found that indole is able to prevent inactivation of tryptophan indole-lyase by trifluoroalanine. The protection of tryptophan indole-lyase by indole exhibits saturation kinetics, with a KD of 0.03 mM, which is comparable to the KI for inhibition of pyruvate ion formation (0.01 mM) and the Km for L-tryptophan synthesis. Fluoride electrode measurements indicate the formation of 28 mol of fluoride ion per mole of enzyme during inactivation of tryptophan indole-lyase, and 121 mol of fluoride ion are formed per mole of enzyme in the presence of 2 mM indole during the same incubation period. 19F NMR spectra of reaction mixtures of tryptophan indole-lyase and trifluoroalanine showed evidence only for fluoride ion formation, in either the absence or the presence of indole, and difluoropyruvic acid was not detected. The partition ratio, kcat/kinact, is estimated to be 9. Tryptophan indole-lyase in the presence of trifluoroalanine exhibits visible absorption peaks at 446 and 478 nm, which decay at the same rate as inactivation. However, in the presence of 1 mM indole and trifluoralanine, tryptophan indole-lyase exhibits a peak only at 420 nm, and the spectra show a gradual increase at 300-310 nm with incubation. In contrast, tryptophan synthase is not protected by indole from inactivation by trifluoroalanine, and the absorption peak at 408 nm for the tryptophan synthase-trifluoroalanine complex is unaffected by indole. These results demonstrate that inactivation of tryptophan indole-lyase occurs via a catalytically competent species, probably the beta,beta-difluoro-alpha-aminoacrylate intermediate, which can be partitioned from inactivation to products by a reactive aromatic nucleophile, indole.  相似文献   

10.
Rabbit liver microsomal preparations fortified with 0.1 mM NADPH effectively promote hydroxylation of [3beta-3H]- or [24-14C]allochenodeoxycholic acid or [5alpha,6alpha-3H2]5alpha-cholestane-3alpha,7alpha-diol to their respective 12alpha-hydroxyl derivatives in yields of about 25 or 65% in 60 min. Minor amounts of other products are formed from the diol. The requirements for activity of rabbit liver microsomal 12alpha-hydroxylase resemble those of rat liver microsomes. Of a number of enzyme inhibitors studied only p-chloromercuribenzoate demonstrated a marked ability to inhibit the reaction with either tritiated substrate. There was no difference in the quantity of product produced from the tritiated acid or the 14C-labeled acid. No clear sex difference was found in activity of the enzyme, nor was an appreciable difference noted in activity of the enzyme between mature and immature animals.  相似文献   

11.
Succinate dehydrogenase (SDH) was solubilized from membranes of Mycobacterium phlei by Triton X-100 with a recovery of about 90%. The solubilized SDH was purified about 90-fold by Sephacryl S-300, DEAE-cellulose, hydroxylapatite, and isoelectric focusing in the presence of Triton X-100 with a 20% recovery. SDH was homogeneous, as determined by polyacrylamide gel electrophoresis in nondenaturing gels containing Triton X-100. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis of the enzyme revealed two subunits with molecular weights of 62,000 and 26,000. SDH is a flavoprotein containing 1 mol of flavin adenine dinucleotide, 7 to 8 mol of nonheme iron, and 7 to 8 mol of acid-labile sulfide per mol of protein. Using phenazine methosulfate and 2,6-dichloroindophenol as electron acceptors, the enzyme had an apparent Km of 0.12 mM succinate. SDH exhibited a sigmoidal relationship of rate to succinate concentration, indicating cooperativity. The enzyme was competitively inhibited by fumarate with a Ki of 0.15 mM. In the absence of Triton X-100, the enzyme aggregated, retained 50% of the activity, and could be resolubilized with Triton X-100 with full restoration of activity. Cardiolipin had no effect on the enzyme activity in the absence of Triton X-100, but it stimulated the activity by about 30% in the presence of 0.1% Triton X-100 in the assay mixture. Menaquinone-9(2H), isolated from M. phlei, had no effect on the enzyme activity either in the presence or absence of Triton X-100.  相似文献   

12.
An enzyme capable of the oxidation of hydroxylamine to nitrite was isolated from the obligate methylotroph Methylococcus capsulatus Bath. The absorption spectra in cell extracts, electron paramagnetic resonance spectra, molecular weight, covalent attachment of heme group to polypeptide, and enzymatic activities suggest that the enzyme is similar to cytochrome P-460, a novel iron-containing protein previously observed only in Nitrosomonas europaea. The native and subunit molecular masses of the M. capsulatus Bath protein were 38,900 and 16,390 Da, respectively; the isoelectric point was 6.98. The enzyme has approximately one iron and one copper atom per subunit. The electron paramagnetic resonance spectrum of the protein showed evidence for a high-spin ferric heme. In contrast to the enzyme from N. europaea, a 13-nm blue shift in the soret band of the ferrocytochrome (463 nm in cell extracts to 450 nm in the final sample) occurred during purification. The amino acid composition and N-terminal amino acid sequence of the enzyme from M. capsulatus Bath was similar but not identical to those of cytochrome P-460 of N. europaea. In cell extracts, the identity of the biological electron acceptor is as yet unestablished. Cytochrome c-555 is able to accept electrons from cytochrome P-460, although the purified enzyme required phenazine methosulfate for maximum hydroxylamine oxidation activity (specific activity, 366 mol of O2 per s per mol of enzyme). Hydroxylamine oxidation rates were stimulated approximately 2-fold by 1 mM cyanide and 1.5-fold by 0.1 mM 8-hydroxyquinoline.  相似文献   

13.
1. Mitochondria isolated from the kidneys of rachitic pigs have been shown to contain an active 25-hydroxyvitamin D3-1 alpha-hydroxylase. From these mitochondria a cytochrome P-450 has been solubilized with a specific content of 0.02-0.04 nmol/mg protein. 2. In the presence of a bovine adrenal NADPH-ferredoxin reductase, bovine adrenal ferredoxin and NADPH, the cytochrome P-450 supported the formation of 1,25-dihydroxyvitamin D3 from 25-hydroxyvitamin D3. 3. The hydroxylation reaction was linear with time up to 40 min, and with the amount of enzyme up to 0.03 nmol cytochrome P-450. The pH optimum for the reaction was 7.4, and the apparent Km was 3 x 10(-10) mol/mg protein. 4. The results show that 25-hydroxyvitamin D3 is metabolized in mammals by the same enzyme system as has been demonstrated in birds.  相似文献   

14.
A basidiomycete, Coprinus sp. SF-1, was found to produce an L-Trp-oxidizing enzyme by screening from the culture collection of our laboratory. After solubilization by 1 M NaSCN from the particulate fraction of disrupted cells of the strain, the enzyme was purified about 76-fold to essential homogeneity. The enzyme had a molecular mass of about 420 kDa and the subunit molecular mass was 68 kDa. The enzyme contained 1 mol of non-covalently bound FAD per mol of the subunit. It catalyzed the simultaneous reactions of oxidative deamination and oxygenative decarboxylation of L-Trp to form indolepyruvic acid and indole-3-acetamide, the former of which was further oxidized to indole-3-acetic acid. The molar ratio of the respective reaction products was about 9:1. The enzyme specifically oxidized L-Trp, and slightly acted on L-Phe and L-Tyr. The Km for L-Trp was about 0.5 mM in both oxidase and oxygenase reactions. Thus, the enzyme is a novel one and was tentatively designated "L-Trp oxidase (deaminating and decarboxylating)". The optimum pHs of oxidase and oxygenase activities were 7.0 and 9.0, respectively. The optimum temperatures of both activities were 50 degrees C. The enzyme was stable at pH 6.0-10.5 and below 50 degrees C, and at 4 degrees C for 1 year.  相似文献   

15.
Previous work showed that random mutagenesis produced a mutant of toluene ortho-monooxygenase (TOM) of Burkholderia cepacia G4 containing the V106A substitution in the hydroxylase -subunit (TomA3) that changed the color of the cell suspension from wild-type brown to green in rich medium. Here, DNA shuffling was used to isolate a random TOM mutant that turned blue due to mutation TomA3 A113V. To better understand the TOM reaction mechanism, we studied the specificity of indole hydroxylation using a spectrum of colored TOM mutants expressed in Escherichia coli TG1 and formed as a result of saturation mutagenesis at TomA3 positions A113 and V106. Colonies expressing these altered enzymes ranged in color from blue through green and purple to orange; and the enzyme products were identified using thin-layer chromatography, high performance liquid chromatography, and liquid chromatography–mass spectroscopy. Derived from the single TOM template, enzymes were identified that produced primarily isoindigo (wild-type TOM), indigo (A113V), indirubin (A113I), and isatin (A113H and V106A/A113G). The discovery that wild-type TOM formed isoindigo via C-2 hydroxylation of the indole pyrrole ring makes this the first oxygenase shown to form this compound. Variant TOM A113G was unable to form indigo, indirubin, or isoindigo (did not hydroxylate the indole pyrrole ring), but produced 4-hydroxyindole and unknown yellow compounds from C-4 hydroxylation of the indole benzene ring. Mutations at V106 in addition to A113G restored C-3 indole oxidation, so along with C-2 indole oxidation, isatin, indigo, and indirubin were formed. Other TomA3 V106/A113 mutants with hydrophobic, polar, or charged amino acids in place of the Val and/or Ala residues hydroxylated indole at the C-3 and C-2 positions, forming isatin, indigo, and indirubin in a variety of distributions. Hence, for the first time, a single enzyme was genetically modified to produce a wide range of colors from indole.  相似文献   

16.
Directed enzyme evolution of 2-hydroxybiphenyl 3-monooxygenase (HbpA; EC ) from Pseudomonas azelaica HBP1 resulted in an enzyme variant (HbpA(ind)) that hydroxylates indole and indole derivatives such as hydroxyindoles and 5-bromoindole. The wild-type protein does not catalyze these reactions. HbpA(ind) contains amino acid substitutions D222V and V368A. The activity for indole hydroxylation was increased 18-fold in this variant. Concomitantly, the K(d) value for indole decreased from 1.5 mm to 78 microm. Investigation of the major reaction products of HbpA(ind) with indole revealed hydroxylation at the carbons of the pyrrole ring of the substrate. Subsequent enzyme-independent condensation and oxidation of the reaction products led to the formation of indigo and indirubin. The activity of the HbpA(ind) mutant monooxygenase for the natural substrate 2-hydroxybiphenyl was six times lower than that of the wild-type enzyme. In HbpA(ind), there was significantly increased uncoupling of NADH oxidation from 2-hydroxybiphenyl hydroxylation, which could be attributed to the substitution D222V. The position of Asp(222) in HbpA, the chemical properties of this residue, and the effects of its substitution indicate that Asp(222) is involved in substrate activation in HbpA.  相似文献   

17.
1. Chromatography on hydroxyapatite and on aminohexyl-Sepharose as well as isoelectric focusing were introduced as new effective purification procedures for phenylalanine racemase (EC 5.1.1.11). The enzyme preparations obtained were essentially homogeneous, as demonstrated by specific activity measurements and polyacrylamide gel electrophoresis. 2. The enzyme is not dissociable by sodium dodecyl sulfate. 3. Phenylalanine racemase is an acidic protein with an isoelectric point of approx. 4.6 (isoelectric focusing). 4. The Michaelis constants of L-Phe and D-Phe in the aminoacyl adenylate activation are 0.06 and 0.13 mM, respectively. 5. From our studies with structural analogues of phenylalanine we infer that the amino group of this amino acid is essential for its binding to the aminoacyl adenylate reaction center. The carboxyl group is not at all or only weakly bound. The benzene ring of phenylalanine which determines substrate recognition also seems to be of minor importance for substrate binding.  相似文献   

18.
A novel enzyme, arylalkyl acylamidase, which shows a strict specificity for N-acetyl arylalkylamines, but not acetanilide derivatives, was purified from the culture broth of Pseudomonas putida Sc2. The purified enzyme appeared to be homogeneous, as judged by native and SDS/PAGE. The enzyme has a molecular mass of approximately 150 kDa and consists of four identical subunits. The purified enzyme catalyzed the hydrolysis of N-acetyl-2-phenylethylamine to 2-phenylethylamine and acetic acid at the rate of 6.25 mumol.min-1.mg-1 at 30 degrees C. It also catalyzed the hydrolysis of various N-acetyl arylalkylamines containing a benzene or indole ring, and acetic acid arylalkyl esters. The enzyme did not hydrolyze acetanilide, N-acetyl aliphatic amines, N-acetyl amino acids, N-acetyl amino sugars or acylthiocholine. The apparent Km for N-acetylbenzylamine, N-acetyl-2-phenylethylamine and N-acetyl-3-phenylpropylamine are 41 mM, 0.31 mM and 1.6 mM, respectively. The purified enzyme was sensitive to thiol reagents such as Ag2SO4, HgCl2 and p-chloromercuribenzoic acid, and its activity was enhanced by divalent metal ions such as Zn2+, Mg2+ and Mn2+.  相似文献   

19.
The mutagenicities of 8 indole compounds (indole-3-acetonitrile, indole-3-carbinol, indole-3-acetamide, indole-3-acetic acid, 3-methylindole, indole-3-aldehyde, indole-3-carboxylic acid and indole) derived from indole glucosinolate were studied by mutation tests on Salmonella typhimurium TA98 and TA100 and Escherichia coli WP2 uvrA/pKM101 with and without S9 mix. None of the 8 indole compounds were mutagenic, but they became mutagenic on these 3 tester strains when treated with nitrite at pH 3. The nitrite-treated indole compounds were mutagenic without metabolic activation system (S9 mix), and their mutagenicities were decreased by the addition of S9 mix.  相似文献   

20.
Degradation of indole and quinoline by Desulfobacterium␣indolicum was studied in batch cultures. The first step in the degradation pathway of indole and quinoline was a hydroxylation at the 2 position to oxindole and 2-hydroxyquinoline respectively. These hydroxylation reactions followed saturation kinetics. The kinetic parameters for indole were an apparent maximum specific transformation rate (V Amax) of 263 μmol mg total protein−1 day−1 and an apparent half-saturation constant (K Am) of 139 μM. The V Amax for quinoline was 170 μmol mg total protein−1 day−1 and K Am was 92 μM. Oxindole inhibited indole hydroxylation whereas 2-hydroxyquinoline stimulated quinoline hydroxylation. An adaptation period of approximately 20 days was required before transformation of 2-hydroxyquinoline in cultures previously grown on quinoline. Indole and quinoline were hydroxylated with a lag phase shorter than 4 h in a culture adapted to ethanol. Chloramphenicol inhibited the hydroxylation of indole and quinoline in ethanol-adapted cells, indicating an inducible enzyme system. Chloramphenicol had no effect on the hydroxylation of indole in quinoline-adapted cells or on the hydroxylation of quinoline in indole-adapted cells. This indicated that it was the same inducible enzyme system that hydroxylated indole and quinoline. Received: 16 July 1996 / Received revision: 23 September 1996 / Accepted: 29 September 1996  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