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 93 毫秒
1.
Processes of aerobic biodegradation of components of phenol production sewage (phenol, acetophenone, dimethylphenylcarbinol, cumene hydroperoxide, α-methylstyrene, benzoate, andp-hydroxybenzoate) by bacterial strains obtained from the collection of the Saratov Institute of Biocatalysis were studied. The metabolic reactions were shown to be oxidative and to have a common catabolic sequence (cumene hydroperoxide-dimethylphenylcarbinol-α-methylstyrene-acetophenone-phenyl acetate-phenol-pyrocatechol-aromatic ring breakage). Benzoate andp-hydroxybenzoate were degraded through the formation of pyrocatechol and protocatechuate, respectively. Metabolic pathways were similar in model mixtures of components and sewage samples.  相似文献   

2.
Rhizobium japonicum 61-A-101 and its bacteroids catabolize phenol and p-hydroxybenzoate. With phenol as a carbon source, utilization started only after a prolonged lag phase while p-hydroxybenzoate was almost instantancously metabolized. Succinate, which supports rapid growth of Rhizobium japonicum, completely repressed respication of phenol; the oxidation of p-hydroxybenzoate was partially inhibited. Pyruvate, supporting slower growth than succinate, retarded the onset of phenol consumption but did not affect its maximum rate.Catabolite repression of phenol utilization by succinate appears to be a characteristic feature of rhizobia. In Pseudomonas putida which also actively metabolizes phenol, succinate had no effect on phenol utilization.  相似文献   

3.
Ubiquinone was biosynthesized when rat liver mitochondria were incubated with S-adenosyl-L-methionine, solanesyl diphosphate, and [U-14C]p-hydroxybenzoate. The intermediates of ubiquinone biosynthesis but not ubiquinone were accumulated in mitochondria incubated without S-adenosyl-L-methionine and the accumulated intermediates were converted to ubiquinone by the addition of the methyl group donor and an excess of cold p-hydroxybenzoate. No solaneylated compounds except nonaprenyl p-hydroxybenzoate were found in sonicated mitochondria, while the biosynthesis of ubiquinone was observed in the sonicated preparation of mitochondria in which the intermediates accumulated. The results indicate that the initial decarboxylation reaction is completely abolished and the subsequent reactions of hydroxylation and methylation are not completely inhibited by the sonication treatment and therefore the decarboxylation reaction is the next step after nonaprenylation of p-hydroxybenzoate. Mitoplasts could biosynthesize ubiquinone with activity comparable to that of intact mitochondria, suggesting that components of the outer membrane and the intermembranous space of mitochondria are not involved in ubiquinone biosynthesis.  相似文献   

4.
Cometabolic degradation of TCE by toluene-degrading bacteria has the potential for being a cost-effective bioremediation technology. However, the application of toluene may pose environmental problems. In this study, several plant essential oils and their components were examined as alternative inducer for TCE cometabolic degradation in a toluene-degrading bacterium, Rhodococcus sp. L4. Using the initial TCE concentration of 80 muM, lemon and lemongrass oil-grown cells were capable of 20 +/- 6% and 27 +/- 8% TCE degradation, which were lower than that of toluene-grown cells (57 +/- 5%). The ability of TCE degradation increased to 36 +/- 6% when the bacterium was induced with cumin oil. The induction of TCE-degrading enzymes was suggested to be due to the presence of citral, cumin aldehyde, cumene, and limonene in these essential oils. In particular, the efficiency of cumin aldehyde and cumene as inducers for TCE cometabolic degradation was similar to toluene. TCE transformation capacities (T (c)) for these induced cells were between 9.4 and 15.1 mug of TCE mg cells(-1), which were similar to the known toluene, phenol, propane or ammonia degraders. Since these plant essential oils are abundant and considered non-toxic to humans, they may be applied to stimulate TCE degradation in the environment.  相似文献   

5.
Cultures of Caulobacter crescentus were found to grow on a variety of aromatic compounds. Degradation of benzoate, p-hydroxybenzoate, and phenol was found to occur via beta-ketoadipate. The induction of degradative enzymes such as benzoate 1,2-dioxygenase, the ring cleavage enzyme catechol 1,2-dioxygenase, and cis, cis-muconate lactonizing enzyme appeared similar to the control mechanism present in Pseudomonas spp. Both benzoate 1,2-dioxygenase and catechol 1,2-dioxygenase had stringent specificities, as revealed by their action toward substituted benzoates and substituted catechols, respectively.  相似文献   

6.
Two strains of Pseudomonas able to grow on phenol or p-nitrophenol (PNP) were isolated from sewage. Pseudomonas sp. PN101 mineralized and formed nitrite from PNP but did not mineralize phenol, and Pseudomonas sp. PH111 mineralized phenol but not PNP. Phenol increased the lag period before Pseudomonas sp. PN101 grew on and mineralized PNP, but this toxicity was reduced by inoculation of the medium with Pseudomonas sp. PH111. PNP inhibited growth of Pseudomonas sp. PH111 and slightly increased the length of the acclimation period for the mineralization of phenol by the bacterium. Inoculation of Pseudomonas sp. PN101 into solutions containing PNP and phenol increased the lag period prior to growth of Pseudomonas sp. PH111 on phenol and markedly lengthened the lag period for its mineralization of phenol. Coinciding with this delay in the onset of phenol degradation was the accumulation of an organic compound formed from PNP by Pseudomonas sp. PN101. This compound was not mineralized by the phenol-degrading bacterium. The data suggest that bacteria may interact during the decomposition of chemical mixtures by destroying or by forming toxins that affect the biodegradation of individual components of those mixtures.  相似文献   

7.
Clostridium thermoaceticum ATCC 39073 converted vanillate to catechol. Although carboxylated aromatic compounds which did not contain methoxyl groups were not by themselves growth supportive, protocatechuate and p-hydroxybenzoate (nonmethoxylated aromatic compounds) were converted to catechol and phenol, respectively, during carbon monoxide-dependent growth. Syringate is not subject to decarboxylation by C. thermoaceticum (Z. Wu, S. L. Daniel, and H. L. Drake, J. Bacteriol. 170:5705-5708, 1988), and sustained growth at the expense of syringate-derived methoxyl groups was dependent on supplemental CO2. In contrast, vanillate was growth supportive in the absence of supplemental CO2, and 14CO2 was the major 14C-labeled product during [carboxyl-14C]vanillate-dependent growth. Furthermore, the decarboxylation of protocatechuate and p-hydroxybenzoate supported methanol- and 1,2,3-trimethoxybenzene-dependent growth (CO2 is required for growth at the expense of these substrates) when supplemental CO2 was depleted from the growth medium, and the decarboxylation of protocatechuate was concomitant with improved cell yields of methanol cultures. These findings demonstrate that (i) C. thermoaceticum is competent in the decarboxylation of certain aromatic compounds and (ii) under certain conditions, decarboxylation may be integrated to the flow of carbon and energy during acetogenesis.  相似文献   

8.
Pseudomonas putida strain S12palB1 was constructed that produces p-hydroxybenzoate from renewable carbon sources via the central metabolite l-tyrosine. P. putida S12palB1 was based on the platform strain P. putida S12TPL3, which has an optimised carbon flux towards l-tyrosine. Phenylalanine ammonia lyase (Pal) was introduced for the conversion of l-tyrosine into p-coumarate, which is further converted into p-hydroxybenzoate by endogenous enzymes. p-Hydroxybenzoate hydroxylase (PobA) was inactivated to prevent the degradation of p-hydroxybenzoate. These modifications resulted in stable accumulation of p-hydroxybenzoate at a yield of 11% (C-molC-mol(-1)) on glucose or on glycerol in shake flask cultures. In a glycerol-limited fed-batch fermentation, a final p-hydroxybenzoate concentration of 12.9mM (1.8gl(-1)) was obtained, at a yield of 8.5% (C-molC-mol(-1)). A 2-fold increase of the specific p-hydroxybenzoate production rate (q(p)) was observed when l-tyrosine was supplied to a steady-state C-limited chemostat culture of P. putida S12palB1. This implied that l-tyrosine availability was the bottleneck for p-hydroxybenzoate production under these conditions. When p-coumarate was added instead, q(p) increased by a factor 4.7, indicating that Pal activity is the limiting factor when sufficient l-tyrosine is available. Thus, two major leads for further improvement of the p-hydroxybenzoate production by P. putida S12palB1 were identified.  相似文献   

9.
10.
Dissimilation of aromatic compounds by Alcaligenes eutrophus   总被引:14,自引:10,他引:4       下载免费PDF全文
The range of aromatic compounds that support the growth of Alcaligenes eutrophus has been determined, and the pathways used for the dissimilation of these substrates have been explored, largely by enzymatic analyses. The beta-ketoadipate pathway operates in the dissimilation of benzoate and p-hydroxybenzoate; the genetisate pathway, in the dissimilation of m-hydroxybenzoate; and the meta cleavage pathway, in the dissimilation of phenol and p-cresol. l-Tryptophan is oxidized via anthranilate; but the metabolic fate of anthranilate was not established. The metabolism of the three stereoisomers of muconic acid was also examined.  相似文献   

11.
p-Hydroxybenzoate hydroxylase (EC 1.14.13.2) from Pseudomonas fluorescens is a NADPH-dependent, FAD-containing monooxygenase catalyzing the hydroxylation of p-hydroxybenzoate to form 3,4-dihydroxybenzoate in the presence of NADPH and molecular oxygen. The mechanism of this three-substrate reaction was investigated in detail at pH 6.6, 4 degrees C, by steady state kinetics, stopped flow spectrophotometry, and equilibrium binding experiments. The initial velocity patterns are consistent with a ping-pong type mechanism which involves two ternary complexes between the enzyme and substrates. The first ternary complex is formed by random addition of p-hydroxybenzoate and NADPH to the enzyme, followed by the release of the first product (NADP+). The reduced enzyme . p-hydroxybenzoate complex now reacts with oxygen, the third substrate, to form the second ternary complex. The enzyme-bound p-hydroxybenzoate then reacts with the activated oxygen to give 3,4-dihydroxybenzoate which is released regenerating the oxidized enzyme for the next cycle. The binding of p-hydroxybenzoate to the oxidized enzyme to form a 1:1 complex causes large, characteristic spectral perturbations and fluorescence quenching. The dissociation constant for the enzyme . substrate complex was obtained by titrations in which absorbance and/or fluorescence quenching was measured. The binding constants of NADPH to the enzyme with and without p-hydroxybenzoate were determined kinetically by measuring the rate of reduction of the enzyme at different concentrations of NADPH. The reduction of the enzyme proceeds extremely slowly in the absence of p-hydroxybenzoate. The presence of the substrate causes a dramatic stimulation (140,000-fold) in the rate of enzyme reduction. The anaerobic reduction of the enzyme by NADPH in the presence of p-hydroxybenzoate produces a transient charge-transfer intermediate. On the basis of the proposed mechanism, the dissociation constants for p-hydroxybenzoate and NADPH as well as the Michaelis constants for all the three substrates were calculated from the initial velocity data. The agreement obtained between various kinetic parameters from the initial rate measurements and those calculated from the individual rate constants determined in rapid reactions, strongly supports the proposed mechanism for the p-hydroxybenzoate hydroxylase reaction.  相似文献   

12.
The inhibition of the proteolytic activity of acrosin in human spermatozoa by butyl p-hydroxybenzoate was assessed by the gelatin substrate film method. Compared with a typical acrosin inhibitor, TLCK, the inhibitory activity of butyl p-hydroxybenzoate to acrosin was much more effective (20 times) than that of TLCK, proving that butyl p-hydroxybenzoate was a potent acrosin inhibitor. The effect of butyl p-hydroxybenzoate on membrane function of human spermatozoa was evaluated using a sperm-tail hypoosmotic swelling test and supravital stain method. A good correlation (r = 0.92) was observed between the % spermatozoa with normal membrane function and the % live spermatozoa after treatment of the spermatozoa with butyl p-hydroxybenzoate for 1 min, indicating that the death of spermatozoa caused by butyl p-hydroxybenzoate is probably due to impairment of sperm membrane function. Both the inhibitory effect on acrosin and the adverse effect on membrane function suggest that butyl p-hydroxybenzoate could be developed as a new vaginal contraceptive.  相似文献   

13.
从哈尔滨太平污水厂活性污泥中筛选到7株高效苯酚降解菌,可利用苯酚作为唯一碳源和能源。通过对这7株菌在不同温度、pH值、以及不同苯酚浓度下生长和苯酚降解情况的考察,确定了这7株菌的最适生长温度为10°C,最适pH值为7.5,最大可降解苯酚浓度为3000mg/L。通过对这7株苯酚降解菌降解性能的研究表明:其具有较强的苯酚降解能力,在10°C、pH值为7.5、装液量为50mL、接种量15%、摇床振荡速度160r/min的条件下,反应48h后可使500mg/L的苯酚降解率达90%以上。葡萄糖对菌体的生长及苯酚降解能力均有一定的影响,当葡萄糖浓度是500mg/L时,该菌对苯酚的降解率仍在80%以上。该研究对处理含有其它碳源的含酚废水具有一定的意义。通过DGGE图谱条带的分析表明,其亮度可以说明这些菌在各个系统中均表现为优势菌,且在污水环境中表现出较强的活性,其优势地位能够稳定地存在。其中2、4、24、28条带丰富,表现出它们在污水环境系统中的多样性。  相似文献   

14.
Chemical Disinfection of Holding-Tank Sewage   总被引:2,自引:2,他引:0       下载免费PDF全文
A number of chemical disinfectants were evaluated for their bactericidal and virucidal effectiveness in holding-tank sewage. It was found that the disinfection efficiencies of formaldehyde, benzalkonium chloride, cetylpyridinium chloride, and methylene blue were markedly improved if the pH of the sewage was raised from 8.0 to 10.5. When formaldehyde, benzalkonium chloride, and methylene blue were tested with either 2-week holding times with no sewage additions or 10-day holding times with daily sewage additions, disinfection effectiveness was maintained as long as the sewage pH was kept at 10.5 and the disinfectant concentration was kept at 100 mg/liter or more. Calcium hypochlorite, zinc sulfate, and phenol were found to be relatively ineffective disinfectants for holding-tank sewage.  相似文献   

15.
A study was conducted to determine the role of concentration of the test chemical, of a second organic compound, and of mutation in the acclimation period before the mineralization of organic compounds in sewage. The acclimation period for the mineralization in sewage of 2 micrograms of 4-nitrophenol (PNP) per liter increased from 6 to 12 days in the presence of 10 mg of 2,4-dinitrophenol per liter. The extension of the acclimation period was equivalent to the time required for mineralization of 2,4-dinitrophenol. In contrast, the time for acclimation for the degradation of 2 micrograms of PNP per liter was reduced when 10 or 100 mg of phenol per liter was added. Lower phenol levels increased the acclimation period to 8 days. The length of the acclimation period for PNP mineralization decreased as the initial concentration of PNP increased from 2 micrograms to 100 mg/liter. The acclimation period for phenol mineralization was lengthened as the phenol concentration increased from 100 to 1,400 mg/liter. The length of the acclimation period for PNP and phenol biodegradation was reproducible, but it varied among replicates for the biodegradation of other nitro-substituted compounds added to sewage or lake water, suggesting that a mutation was responsible for acclimation to these other compounds. The acclimation period may thus reflect the time required for the destruction of toxins, and it also may be affected by the concentration of the test compound or the presence of other substrates.  相似文献   

16.
A study was conducted to determine the role of concentration of the test chemical, of a second organic compound, and of mutation in the acclimation period before the mineralization of organic compounds in sewage. The acclimation period for the mineralization in sewage of 2 micrograms of 4-nitrophenol (PNP) per liter increased from 6 to 12 days in the presence of 10 mg of 2,4-dinitrophenol per liter. The extension of the acclimation period was equivalent to the time required for mineralization of 2,4-dinitrophenol. In contrast, the time for acclimation for the degradation of 2 micrograms of PNP per liter was reduced when 10 or 100 mg of phenol per liter was added. Lower phenol levels increased the acclimation period to 8 days. The length of the acclimation period for PNP mineralization decreased as the initial concentration of PNP increased from 2 micrograms to 100 mg/liter. The acclimation period for phenol mineralization was lengthened as the phenol concentration increased from 100 to 1,400 mg/liter. The length of the acclimation period for PNP and phenol biodegradation was reproducible, but it varied among replicates for the biodegradation of other nitro-substituted compounds added to sewage or lake water, suggesting that a mutation was responsible for acclimation to these other compounds. The acclimation period may thus reflect the time required for the destruction of toxins, and it also may be affected by the concentration of the test compound or the presence of other substrates.  相似文献   

17.
Aerobic degradation of 7 mmol/L phenol in the presence of alternative carbon sources (7 mmol/L glucose or acetate or 1–2 mmol/L 2‐chlorophenol) was investigated using non‐acclimatized and acclimatized sewage sludges and enrichment cultures. The substrates represented an intermediate of phenol degradation (acetate), an independent substrate (glucose) or a “precursor‐substrate” of phenol degradation (2‐chlorophenol). Bacteria from sewage sludge, not pre‐adapted to phenol (2 mmol/L), rapidly respired acetate and glucose in the presence of phenol, whereas phenol was only bioconverted to any unknown aromatic metabolite after 24 h. In the presence of phenol and 2‐chlorophenol, no removal of both substances was observed when using the unacclimatized sludge. Sludge that was acclimatized to the degradation of phenol showed an initial preference for easily degradable co‐substrates such as glucose or acetate with only a slow concomitant respiration of phenol. Respiration of phenol increased rapidly after the co‐substrates were depleted. The highest phenol degradation rates were 51.6 mmol/L d, when phenol was the sole carbon substrate. Vice versa, phenol was preferentially respired in the presence of a less easily degradable co‐substrate such as 2‐chlorophenol at a rate of around 7 mmol/L d. Further studies with an enrichment culture that was obtained after 7 successive transfers of phenol‐adapted sludge into mineral medium with phenol as the only carbon source indicated that the acetate and glucose‐degrading capabilities were diminished or almost completely lost. In these enrichment cultures, phenol degradation was not affected by the presence of glucose, but glucose was not degraded. In contrary, the presence of acetate slightly slowed down the phenol degradation rate of the enrichment culture. Growth of the microorganisms apparently occurred at the expense of phenol and acetate respiration. The result of this work may be of practical importance in determining the feeding strategy, which is the key factor for most biological wastewater treatment systems. When acetate was present together with phenol in a wastewater, the phenol degradation rates were influenced by acetate, since acetate was an intermediate of phenol degradation. Glucose as an “independent substrate” was apparently degraded by other bacteria via acetate, and in this way it also influenced the phenol degradation rates. Glucose‐degrading bacteria could be “washed out” from the acclimatized sludge during several transfers into mineral medium with phenol as the sole carbon source. If later on, glucose was added again, it remained undegraded and did not influence phenol degradation. 2‐Chlorophenol degradation also requires other bacteria than phenol degraders.  相似文献   

18.
Summary Anaerobic phenol degrading consortia were selected in sewage sludge and culture conditions were improved to allow maximum degradation rates of 0.9 g/l·d. Phenol had to be added in two portions of 0.45 g/l at intervals of 12 h to keep the fermentation at stable conditions. From U-14C-phenol little benzoate and acetate were formed as intermediates under a N2:CO2 gas phase. Final products were methane and CO2. When methanogenesis was inhibited by BESA, less labeled methane and CO2 were formed and labeled acetate remained undegraded. Turnover rates of phenol were significantly reduced in the presence of a H2:CO2 gas atmosphere and benzoate was formed from phenol and CO2. Acetate did not accumulate remarkably. After the H2:CO2 was converted to methane or was exchanged by N2:CO2 the accumulated benzoate was further degraded to methane and CO2. Elevated pools of acetate in sewage sludge led also to a reduction of the phenol degradation rates and presumably to an increased concentration of benzoate. In fresh sewage sludge benzoate degradation proceeds immediately, while the degradation of phenol starts only after a lag-phase of 3–10 days.  相似文献   

19.
The flavin prosthetic group (FAD) of p-hydroxybenzoate hydroxylase from Pseudomonas fluorescens was replaced by a stereochemical analog, which is spontaneously formed from natural FAD in alcohol oxidases from methylotrophic yeasts. Reconstitution of p-hydroxybenzoate hydroxylase from apoprotein and modified FAD is a rapid process complete within seconds. Crystals of the enzyme-substrate complex of modified FAD-containing p-hydroxybenzoate hydroxylase diffract to 2.1 A resolution. The crystal structure provides direct evidence for the presence of an arabityl sugar chain in the modified form of FAD. The isoalloxazine ring of the arabinoflavin adenine dinucleotide (a-FAD) is located in a cleft outside the active site as recently observed in several other p-hydroxybenzoate hydroxylase complexes. Like the native enzyme, a-FAD-containing p-hydroxybenzoate hydroxylase preferentially binds the phenolate form of the substrate (pKo = 7.2). The substrate acts as an effector highly stimulating the rate of enzyme reduction by NADPH (kred > 500 s-1). The oxidative part of the catalytic cycle of a-FAD-containing p-hydroxybenzoate hydroxylase differs from native enzyme. Partial uncoupling of hydroxylation results in the formation of about 0.3 mol of 3,4-dihydroxybenzoate and 0.7 mol of hydrogen peroxide per mol NADPH oxidized. It is proposed that flavin motion in p-hydroxybenzoate hydroxylase is important for efficient reduction and that the flavin "out" conformation is associated with the oxidase activity.  相似文献   

20.
Biodegradation of phenol by a mixed microbial culture, isolated from a sewage treatment plant, was investigated in batch shake flasks. A minimum concentration of 100 and a maximum of 800 mg 1(-1) of phenol in the media were adapted in the degradation study. The phenol degradation rate varied largely and was less than 10 mg l(-1)h(-1) at both extremes of the initial concentrations in the media. The degradation rate was maximum 15.7 mg l(-1)h(-1) at 400 mg l(-1) phenol. The culture followed substrate inhibition kinetics and the specific growth rate were fitted to Haldane and Han-Levenspiel models. Between the two models the Han-Levenspiel was found to be a better fit with a root mean square error of 0.0211. The biokinetics constants estimated using these models showed good potential of the mixed microbial culture in phenol degradation.  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