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 15 毫秒
1.
Eight lots of reagent-grade phenol from four companies were tested for capacity to interact with Cu2+ to produce an inactivator or inactivators of the transfective RNA obtained from poliovirions; such capacity to interact with Cu2+ is referred to as cofactor activity. Six of the lots showed cofactor activity; two did not. A review of the data on the phenol lots and of the properties of the impurity or impurities conferring cofactor activity suggested that the active impurity(ies) might be a dihydric or trihydric phenol. Commercial catechol, resorcinol, hydroquinone, orcinol and pyrogallol were tested and found active. The activity of hydroquinone was outstandingly high. Upon serial recrystallization, the activity of catechol, hydroquinone, orcinol and pyrogallol remained constant, but the activity of resorcinol decreased markedly, in stepwise fashion, showing the most of the activity of the commercial resorcinol was due to impurity(ies). Each of catechol, hydroquinone, orcinol, pyrogallol, and the commercial resorcinol was shown to react with Cu2+ to produce inactivator(s). The effective target for inactivator(s) was the RNA and not the transfection process. The kinetics of inactivator(s) production varied for the different phenols, and the inactivator activity of the incubated mixture of pyrogallol and Cu2+ was notably labile.  相似文献   

2.
J K Lin  S F Lee 《Mutation research》1992,269(2):217-224
The hydrolytic products of lignins, humic acids and industrial waste including hydroquinone, catechol, resorcinol, pyrogallol and 1,2,4-benzenetriol are widely distributed in water sources. These polyphenols can interact with chlorine or nitrite to yield new derivatives. Generally, these new products possess more mutagenic potential than their original compounds. Furthermore, the mutagenicity of these polyphenols and their derivatives can be dramatically reduced by rodent liver microsomal enzymes (S9). The mutagenicity of polyphenols is in this order: hydroquinone greater than 1,2,4-benzenetriol greater than pyrogallol, while catechol, resorcinol and phloroglucinol are non-mutagenic. The ultimate product of chlorination or nitrosation of hydroquinone has been identified to be p-benzoquinone. The formation of active oxygen species including superoxide anion and hydrogen peroxide by polyphenols has been demonstrated and this may contribute partly to the molecular mechanisms of polyphenol mutagenicity.  相似文献   

3.
The pathway for the aerobic catabolism of 1,3,5-trihydroxybenzene (phloroglucinol) by a new strain of Penicillium was investigated using both in vivo and in vitro cell-free systems. The fungal strain was isolated by enrichment on phloroglucinol and identified as P. simplicissimum (Oud) Thom. It grew optimally at pH 5.5 and 27°C with 119 mM (1.5%w/v) of phloroglucinol in a basal mineral salts medium. Vapours of the crystalline substrate placed in a Petri-plate lid supported the growth of the fungal colonies on the agar surface. Mycelia grown on phloroglucinol accumulated 1,2,4-trihydroxybenzene and resorcinol in the medium. Washed, resting mycelia grown on phloroglucinol, when resuspended in a buffer utilized oxygen in the presence of catechol, resorcinol, pyrogallol and phloroglucinol. A NADPH-dependent reductase in the cell-free extract reduced phloroglucinol to dihydrophloroglucinol. This electron donor could not be replaced by NADH. Resorcinol hydroxylase, phloroglucinol reductase, catechol-1,2-oxygenase, and catechol-2,3-oxygenase were detected in cell-free extracts of mycelia grown on phloroglucinol. The possible steps in the degradation of phloroglucinol are discussed.  相似文献   

4.
Benzene is strongly suspected of being an animal and human carcinogen, but the mechanisms by which it induces tumors of lymphoid and hematopoietic organs are unknown. Production of active oxygen species from benzene metabolites [hydroquinone (HQ), catechol and 1,2,4-benzenetriol (1,2,4-BT) and related polyphenols (resorcinol, pyrogallol and phloroglucinol) are investigated. Pyrogallol and 1,2,4-BT can produce H2O2, O 2 and·OH simultaneously, and have powerful mutagenic potential. Resorcinol and phloroglucinol cannot produce all of the active oxygen species, and show no mutagenic effects. Catechol can produce H2O2, but cannot produce O 2 and·OH, and has no mutagenic activity. These data strongly support the hypothesis that benzene metabolites can cause mutagenicity via the generation of oxygen radicals. Although HQ produces H2O2 only, and less than produced by pyrogallol and 1,2,4-BT, the mutagenicity of HQ is higher. The results indicate that HQ may act via another mechanism to cause mutagenicity. In the presence of trace metal ions, the reactivity of polyphenols is increased. The biological significance of these phenomena are investigated and discussed.  相似文献   

5.
Summary The effect of phenolic compounds in foodstuffs on histamine and leukotriene B4 (LTB4) release from rat peritoneal exudate cells and their antioxidative activity were examined to assess their antiallergenic activities. Among them, triphenols such as pyrogallol and gallic acid inhibited histamine release from the cells, but diphenols did not. On the other hand, o- and p-diphenols such as catechol and hydroquinone with strong antioxidative activity inhibited LTB4 release as strongly as pyrogallol, but an m-derivative resorcinol with weak antioxidative activity did not. Though carboxylated compounds and their noncarboxylated counterparts were antioxidative, the former exerted a much weaker inhibitory effect on the LTB4 release than the latter. In flavonols, only myricetin with a triphenolic B ring strongly inhibited histamine release, but all flavonols strongly suppressed LTB4 release irrespective of the number of OH groups in the B ring. Among flavonoids with an o-diphenolic B ring, flavonol and flavone with a C4-carbonyl group strongly inhibited LTB4 release, whereas the activity of anthocyan without C4-carbonyl was much weaker than the above compounds. These results suggest that triphenolic structure is essential for the inhibition of histamine release. On the other hand, antioxidative activity and membrane permeability of phenolic compounds seemed to be essential for the inhibition of LTB4 release. In addition, the C4-carbonyl group seemed to be important for strongly inhibiting LTB4 release.  相似文献   

6.
Ibuprofen, one of the nonsteroidal anti-inflammatory drugs, inhibited arylamine N-acetyltransferase activity of Klebsiella pneumoniae both in vitro and in vivo. The NAT activities of Klebsiella pneumoniae were inhibited by ibuprofen in a dose-dependent manner both in vitro and in vivo. In vitro, the NAT activity was 0.675 ± 0.028 nmol/min/mg of protein for the acetylation of 2-aminofluorene. In the presence of 8 mM ibuprofen, the NAT activity was 0.506 ± 0.002 nmol/min/mg of protein for the acetylation of 2-aminofluorene. In vivo, the NAT activity was 0.279 ± 0.016 nmol/min/1010 colony forming units (CFU) for the acetylation of 2-aminofluorene. In the presence of 8 mM ibuprofen, the NAT activity was 0.228 ± 0.008 nmol/min/1010 CFU for the acetylation of 2-aminofluorene. The inhibition of NAT activity by ibuprofen was shown to persist for at least 4 h. For in vitro examination, the values of apparent K m and V max were 1.08 ± 0.05 mM and 9.17 ± 0.11 nmol/min/mg of protein, respectively, for 2-aminofluorene. However, when 8 mM of ibuprofen was added to the reaction mixtures, the values of apparent K m and V max were 1.19 ± 0.01 mM and 6.67 ± 0.11 nmol/min/mg of protein, respectively, for 2-aminofluorene. For in vivo examination, the values of apparent K m and V max were 1.24 ± 0.48 mM and 4.18 ± 1.06 nmol/min/10 × 1010 CFU, respectively, for 2-aminofluorene. However, when 8 mM of ibuprofen was added to the culture, the values of apparent K m and V max were 0.95 ± 0.29 mM and 2.77 ± 0.37 nmol/min/mg protein, respectively, for 2-aminofluorene, respectively. This report is the first finding of ibuprofen inhibition of arylamine N-acetyltransferase activity in a strain of Klebsiella pneumoniae. Received: 28 January 1997 / Accepted: 12 February 1997  相似文献   

7.
Trihydroxybenzenes are degraded anaerobically through the phloroglucinol pathway. In Pelobacter acidigallici as well as in Pelobacter massiliensis, pyrogallol is converted to phloroglucinol in the presence of 1,2,3,5-tetrahydroxybenzene by intermolecular hydroxyl transfer. The enzyme catalyzing this reaction was purified to chromatographic and electrophoretic homogeneity. Gel filtration and electrophoresis revealed a heterodimer structure with an apparent molecular mass of 127 kDa for the native enzyme and 86 kDa and 38 kDa, respectively, for the subunits. The enzyme was not sensitive to oxygen. HgCl2, p-chloromercuribenzoic acid, and CuCl2 inhibited strongly the reaction indicating an essential function of SH-groups. Transhydroxylase had a pH-optimum of 7.0 and a pI of 4.1. The apparent temperature optimum was in the range of 53°C to 58°C. The activation energy for the conversion of pyrogallol and 1,2,3,5-tetrahydroxybenzene to phloroglucinol and tetrahydroxybenzene was 31.4 kJ per mol. Purified enzyme exhibited a specific activity of 3.1 mol. m−1 mg−1 protein and an apparent Km for pyrogallol and 1,2,3,5-tetrahydroxybenzene of 0.70 mM and 0.71 mM, respectively. The enzyme was found to contain per mol heterodimer 1.1 mol molybdenum, 12.1 mol iron and 14.5 mol acid-labile sulfur. Requirement for molybdenum for transhydroxylating enzyme activity was proven also by cultivation experiments. No hints for the presence of flavins were obtained. The results presented here support the hypothesis that a redox reaction is involved in this intermolecular hydroxyl transfer.  相似文献   

8.
Heme‐containing catalases and catalase‐peroxidases catalyze the dismutation of hydrogen peroxide as their predominant catalytic activity, but in addition, individual enzymes support low levels of peroxidase and oxidase activities, produce superoxide, and activate isoniazid as an antitubercular drug. The recent report of a heme enzyme with catalase, peroxidase and penicillin oxidase activities in Bacillus pumilus and its categorization as an unusual catalase‐peroxidase led us to investigate the enzyme for comparison with other catalase‐peroxidases, catalases, and peroxidases. Characterization revealed a typical homotetrameric catalase with one pentacoordinated heme b per subunit (Tyr340 being the axial ligand), albeit in two orientations, and a very fast catalatic turnover rate (kcat = 339,000 s?1). In addition, the enzyme supported a much slower (kcat = 20 s?1) peroxidatic activity utilizing substrates as diverse as ABTS and polyphenols, but no oxidase activity. Two binding sites, one in the main access channel and the other on the protein surface, accommodating pyrogallol, catechol, resorcinol, guaiacol, hydroquinone, and 2‐chlorophenol were identified in crystal structures at 1.65–1.95 Å. A third site, in the heme distal side, accommodating only pyrogallol and catechol, interacting with the heme iron and the catalytic His and Arg residues, was also identified. This site was confirmed in solution by EPR spectroscopy characterization, which also showed that the phenolic oxygen was not directly coordinated to the heme iron (no low‐spin conversion of the FeIII high‐spin EPR signal upon substrate binding). This is the first demonstration of phenolic substrates directly accessing the heme distal side of a catalase. Proteins 2015; 83:853–866. © 2015 Wiley Periodicals, Inc.  相似文献   

9.
Hydroquinone, a metabolite of benzene, is converted by human myeloperoxidase to 1,4-benzoquinone, a highly toxic species. This conversion is stimulated by phenol, another metabolite of benzene. Here we report that peroxidase-dependent hydroquinone metabolism is also stimulated by catechol, resorcinol, o-cresol, m-cresol, p-cresol, guaiacol, histidine, and imidazole. In order to gain insights into the mechanisms of this stimulation, we have compared the kinetics of human myeloperoxidase-dependent phenol, hydroquinone, and catechol metabolism. The specificity (Vmax/Km) of hydroquinone for myeloperoxidase was found to be 5-fold greater than that of catechol and 16-fold greater than that of phenol. These specificities for myeloperoxidase-dependent metabolism inversely correlated with the respective one-electron oxidation potentials of hydroquinone, catechol, and phenol and suggested that phenol- and catechol-induced stimulation of myeloperoxidase-dependent hydroquinone metabolism cannot simply be explained by interaction of hydroquinone with stimulant-derived radicals. Phenol (100 microM), catechol (20 microM), and imidazole (50 mM) did, however, all increase the specificity (Vmax/Km) of hydroquinone for myeloperoxidase, indicating that these three compounds may be stimulating hydroquinone metabolism by a common mechanism. Interestingly, the stimulation of peroxidase-dependent hydroquinone metabolism by other phenolic compounds was pH-dependent, with the stimulating effect being higher under alkaline conditions. These results therefore suggest that the interaction of phenolic compounds, presumably by hydrogen-bonding, with the activity limiting distal amino acid residue(s) or with the ferryl oxygen of peroxidase may be an important contributing factor in the enhanced myeloperoxidase-dependent metabolism of hydroquinone in the presence of other phenolic compounds.  相似文献   

10.
The subunit MW of Dioscorea bulbifera polyphenol oxidase (MW 115 000 ± 2000) determined by SDS-PAGE is ca. 31 000 indicating that the enzyme is an oligomeric protein with four subunits. Ki values of various inhibitors and their modes of inhibition have been determined with catechol and pyrogallol as substrates. p-Nitrophenol, p-cresol, quinoline and resorcinol are competitive inhibitors of catechol binding while only orcinol and p-nitrophenol behave in the same way towards pyrogallol as substrate. From the effect of pH on Vmax, groups with pK values ca. 4.7 and 6.8 have been identified to be involved in catalytic activity. The Arrhenius activation energy (Ea) at pH 4.0 is 8.9 kcal/mol between 40–65°. At pH 7.0, the value is 22.1 kcal/mol between 40 and 60°. The enthalpies (ΔH) at pH 4.0 and pH 7.0 are 2.3 kcal/mol and 32.4 kcal/mol respectively. The results are discussed considering the conformational changes of the enzyme during substrate binding.  相似文献   

11.
—Samples of South Vietnamese soils intensely treated with Agent Orange defoliant were tested for the presence of fungi and actinomycetes with an elevated phenol oxidase activity. As a result, a fast-growing nonsporulating strain producing neutral phenol oxidases was isolated and identified asMycelia sterilia INBI2-26. The strain formed extracellular phenol oxidases during surface growth on a liquid medium in the presence of guayacol and copper sulfate, as well as during submerged cultivation in liquid medium containing wheat bran and sugar beet pulp. Isoelectric focusing of the culture liquid revealed two major catechol oxidases (PO1 and PO2) with pI 3.5 and 8, respectively. The enzymes were purified by Ultrafiltration, ion exchange chromatography, and exclusion HPLC. Both were stable between pH 3 and 8. At pH 8 and 40°C., they retained at least 50% of activity after incubation for 50 h. At 50°C., PO2 was more stable and retained 40% of activity after 50 h, whereas PO1 was inactivated in 3–6 h. The pH-optimutns for PO1 and PO2 toward catechol were 6 and 6.5; and theK m values were 1.5±0.35 and 1.25±0.2 mM, respectively. PO1 and PO2 most optimally oxidized 2,2′-azinobis-(3-ethylbenzthiazoline-6-sulfonic acid) at pH 3 withK m values 1.6±0.18 and 0.045±0.01 mM, respectively, but displayed no activity toward tyrosine. The PO2 absorbance spectrum had a peak at 600 nm, thus indicating the enzyme to be a member of the laccase family.  相似文献   

12.
Some aspects of the induction of enzymes participating in the metabolism of phenol and resorcinol in Trichosporon cutaneum were studied using intact cells and cell-free preparations.Activities of phenol hydroxylase (1.14.13.7), catechol 1,2-oxygenase (1.13.11.1), cis,cis-muconate cyclase (5.5.1.-), delactonizing enzyme(s) and maleolylacetate reductase were 50–400 times higher in fully induced cells than in noninduced cells.In addition to phenol and resorcinol, also catechol, cresols and fluorophenols could induce phenol hydroxylase.The induction was severely inhibited by phenol concentrations higher than 1 mM. Using optimum inducer concentrations (0.01–0.10 mM), it took more than 8 h to obtain full induction, whether in proliferating or in nonproliferating cells.Phenol hydroxylase, catechol 1,2-oxygenase and cis,cis-muconate cyclase were induced simultaneously. The synthesis of the de-lactonizing activity was delayed in relation to these three preceeding enzymes of the pathway.High glucose concentration (over 15 mM) inhibited completely the induction of phenol oxidation by nonproliferating cells. It also inhibited phenol oxidation by pre-induced cells.Among the NADPH-generating enzymes, the activity of iso-citrate dehydrogenase was elevated in cells grown on phenol and resorcinol instead of glucose.  相似文献   

13.
Polyphenol oxidase (PPO, EC 1.14.18.1) was extracted from celery roots (Apium graveolens L.) with 0.1 M phosphate buffer, pH 7.0. The PPO was partially purified by (NH4)2SO4 and dialysis. Substrate specificity experiments were carried out with catechol, pyrogallol, L-DOPA, p-cresol, resorcinol, and tyrosine. The Km for pyrogallol, catechol, and L-DOPA were 4.5, 8.3, and 6.2mM, respectively, at 25 degrees C. Data for Vmax/Km values, which represent catalytic efficiency, show that pyrogallol has the highest value. The optimum pH and temperature were determined with catechol, pyrogallol, and L-DOPA. Optimum pH was 7.0 for catechol and L-DOPA, and 7.5 for pyrogallol. Optimum temperatures for maximum PPO activity were 25 degrees C for pyrogallol, 40 degrees C for catechol, and 45 degrees C for L-DOPA. Heat inactivation studies showed a decrease in enzymatic activity at temperatures above 60 degrees C. The order of inhibitor effectiveness was: L-cysteine > ascorbic acid > glycine > resorcinol > NaCl.  相似文献   

14.
Biodegradation of resorcinol, catechol and hydroquinone under anaerobic conditions was studied in identical upflow fixed film-fixed bed reactors. Kinetic constants; V(max) (maximum substrate utilization rate) and K(s) (Monod's half saturation constant) were determined for the three compounds using Lineweaver-Burk plot. V(max) for resorcinol was highest, followed by catechol and then by hydroquinone. When both resorcinol and catechol were fed to the resorcinol-acclimated reactor, resorcinol degradation was inhibited by catechol. The inhibition was of the uncompetitive type and V(max) for resorcinol was reduced by catechol.  相似文献   

15.
In the current study, a dimeric phenoloxidase (PO) from the hemolymph of healthy and diseased (pebrine infected) larvae of Antheraea assamensis Helfer was extracted and purified. The protein was subjected to purification using Sephacryl S‐100 and CM Sepharose chromatography. The enzyme comprised of two subunits of ~76.8 and 76 kDa that showed PO activity in 6 mM l ‐3,4‐dihydroxyphenylalanine (L ‐DOPA) and 8 mM catechol but not in hydroquinone. Optimum temperature for PO activity was 30°C in l ‐DOPA and 37°C in catechol. Optimum pH ranged from 6.8 to 7.0 in L ‐DOPA and 7.0–7.2 in catechol. Specific activity of the purified PO from healthy larvae was 53.9 µM/min per mg of protein per ml in L ‐DOPA and 50.77 µM/min per mg of protein per ml in catechol. Specific activity of PO from diseased larvae was 30.0 µM/min per mg of protein per ml in L ‐DOPA and 28.55 µM/min per mg of protein per ml in catechol. Purification fold was 3.27–4.21 for healthy and 2.38–2.56 for diseased fractions. The enzyme showed the Michaelis constant (Km) of 2.46–2.85 mM for healthy and diseased fractions in L ‐DOPA. In catechol Km of 9.23–17.71 mM was observed. Peptidoglycan was the best activator of purified PO from both healthy and diseased fractions. Interactions between controls and activators appeared statistically significant (F = 767.5; df = 3; P < 0.0001). Na+, K+, and Cu2+ increased, whereas Ca2+, Zn2+, Mg2+, and Co2+ decreased PO activity. The overall interactions appeared highly significant (F = 217.0; df = 27; P < 0.0001). Kojic acid, dithiothreitol, thiourea, phenylthiourea, carbendazim, N‐bromosuccinimide, N,N,N′,N′‐tetraacetic acid, and diethyldithiocarbamate inhibited PO activity.  相似文献   

16.
THE BIOLOGICAL OXIDATION OF SPENT GAS LIQUOR   总被引:4,自引:4,他引:0  
SUMMARY: Mixed cultures of bacteria grown in spent gas liquor readily oxidized phenol, o -, m - and p -cresol, catechol, 3-methyl catechol, 4-methyl catechol, resorcinol, 2-methyl resorcinol, and 4-methyl resorcinol. Quinol, pyrogallol and phloroglucinol were more resistant. The optimum temperature was 30° and the best pH range 6·5–7·8. Yeast extract and sterile sewage sludge both increased the rate of growth of organisms in liquor when the inoculum was small. Five phenol oxidizing organisms were isolated in pure culture. Copper in concentrations greater than 1 p/m inhibited both growth and phenol oxidation by one of these.
Mixed cultures grown in an ammonium thiocyanate medium originally inoculated with Thiobacillus thiocyanoxidans oxidized potassium thiocyanate and sodium thiosulphate. Chloride inhibited thiocyanate oxidation in concentrations above 5,000 p/m, although adaptation to 15,000 p/m was possible. Phenol inhibited thiocyanate oxidation in concentrations of 300 p/m or more. Mixed cultures grown on sodium thiosulphate oxidized sodium trithionate and tetrathionate, potassium pentathionate and hexa-thionate, and potassium and ammonium thiocyanate
Manometric determinations of the 5 day biological oxygen demand of effluents after treatment showed good agreement with the values obtained by the conventional method, the manometric values being usually somewhat higher.  相似文献   

17.
In Neurospora crassa the aliphatic alcohols methanol, ethanol, 1-propanol, 2-propanol, 1-butanol, 2-butanol, isobutanol, ethylene glycol, glycerol, and allyl alcohol and the phenolic compounds phenol, hydroquinone, resorcinol, pyrogallol, phloroglucinol, sodium salicylate, and acetylsalicylic acid were analyzed with respect to their capacities to induce heat shock proteins (HSP) and to inhibit protein synthesis. Both the alcohols and phenols showed the greatest levels of HSP induction at concentrations which inhibited the overall protein synthesis by about 50%. The abilities of the different alcohols to induce the heat shock response are proportional to their lipophilicities: the lipophilic alcohol isobutanol is maximally inductive at about 0.6 M, whereas the least lipophilic alcohol, methanol, causes maximal induction at 5.7 M. The phenols, in general, show a higher capability to induce the heat shock response. The concentrations for maximal induction range between 25 mM (sodium salicylate) and 100 mM (resorcinol). Glycerol (4.1 M) shifted the concentration necessary for maximal HSP induction by hydroquinone from 50 to 200 mM. The results reveal that the induction of HSP occurs under conditions which considerably constrain cell metabolism. The heat shock response, therefore, does not represent a sensitive marker for toxicity tests but provides a good estimate for the extent of cell damage.  相似文献   

18.
In this work, the biodegradation mechanism of phenol and sub products (such as catechol and hydroquinone) in Chromobacterium violaceum was investigated by cloning and molecular characterization of a phenol monooxygenase gene in Escherichia coli. This gene (Cvmp) is very similar (74 and 59% of similarity and identity, respectively) to the ortholog from Ralstonia eutropha bacteria capable of utilizing phenol as the sole carbon source. The phenol biodegradation ability of E. coli recombinant strains was tested by cell-growth in a minimal medium containing phenol as the sole source of carbon and release of intermediary metabolites (catechol and hydroquinone). Interestingly, during the growth of these strains on phenol, catechol, and hydroquinone accumulated transiently in the medium. These metabolites were further analyzed by HPLC. These results indicated that phenol can be initially orto or para hydroxylated to produce cathecol or hydroquinone, respectively, followed by meta-cleavage of aromatic rings. To verify this information, the metabolites obtained from HPLC were submitted to LC/MS to confirm their chemical structure, thereby indicating that the recombinant strains utilize two different routes simultaneously, leading to different ring-fission substrates for the metabolism of phenol.  相似文献   

19.
Crocus sativus L., cultivated since ancient times as the source of saffron, is a triploid plant that can be propagated only via its corms which undergo a period of dormancy. Understanding the processes taking place in the corm is essential to preserve the plant and improve its quality. Color and taste being of prime importance in the quality of the saffron spice, knowledge on polyphenol oxidase (PPO) activity in the plant is of particular interest given the role of the enzyme in fruit and vegetable browning during processing and during the storage of processed food. In this paper, PPO activity was investigated for the first time in extracts obtained from dormant C. sativus L. corms. PPO activity was detectable using l-DOPA, pyrogallol, catechol or p-cresol as substrate, each being oxidized to its corresponding o-quinone; no activity was detectable with l-tyrosine, tyramine or phenol as substrate. Two pH optima, respectively at 4.5 and 6.7, were observed with all substrates and a third one, at 8.5, was found with l-DOPA and p-cresol. Kinetics parameters studied at pH 6.7 indicated the highest catalytic efficiency (in units mg−1 prot mM−1) with pyrogallol: 150, then catechol: 39, l-DOPA: 6.4 and p-cresol: 4.6. The enzymatic activity was inhibited by 50% in the presence of 0.22, 0.35, 0.5 and 0.7 mM kojic acid with, respectively, catechol, pyrogallol, p-cresol and l-DOPA as substrate. When stained for PPO activity, non-denaturing gel electropherograms of extract revealed three distinct bands, indicating the presence of multiple isoenzymes in dormant C. sativus L. corms.  相似文献   

20.
Oxidation of phenols by cells and cell-free enzymes from Candida tropicalis   总被引:5,自引:0,他引:5  
A yeast strain isolated from soil by enrichment on phenol as major carbon source was identified as Candida tropicalis. Washed cell suspensions of this strain and cell-free preparations obtained from mechanically disrupted cells oxidized phenol via catechol and cis, cis-muconate. In addition to phenol and the three isomeric diphenols, a number of phenol derivatives, amongst them fluoro-, nitro- and short-chain alkyl-phenols, were oxidized by the organism. However, no significant oxygen uptake could be demonstrated in the presence of pyrogallol, phloroglucinol, the cresols, the m-and p-hydroxy-benzoates, methoxylated phenol derivatives, benzene or toluene. Cell-free preparations from the yeast strain exhibited activity of phenol hydroxylase and of catechol 1,2-oxygenase. Both enzymes appeared in the soluble cell fraction. Both exhibit broad substrate specificities. The relative specific activity of the ring-cleaving enzyme towards various substrates seems to be dependent on the phenolic inducer.  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