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.
The feasibility of anaerobic methanethiol (MT) degradation at elevated sodium concentrations was investigated in a mesophilic (30 degrees C) lab-scale upflow anaerobic sludge bed (UASB) reactor, inoculated with estuarine sediment originating from the Wadden Sea (The Netherlands). MT was almost completely degraded (>95%) to sulfide, methane and carbon dioxide at volumetric loading rates up to 37 mmol MT x L(-1) x day(-1), 0.5 M sodium (NaCl or NaHCO(3)) and between pH 7.3 and 8.4. Batch experiments revealed that inhibition of MT degradation started at sodium (both NaCl and NaHCO(3)) concentrations exceeding 0.8 M. Sulfide inhibited MT degradation already around 3 mM (pH 8.3).  相似文献   

2.
Anaerobic methanethiol (MT) degradation by mesophilic (30 degrees C) alkaliphilic (pH 10) communities was studied in a lab-scale Upflow Anaerobic Sludge Bed (UASB) reactor inoculated with a mixture of sediments from the Wadden Sea (The Netherlands), Soap Lake (Central Washington), and Russian soda lakes. MT degradation started after 32 days of incubation. During the first 252 days, complete degradation was achieved till a volumetric loading rate of 7.5 mmol MT/L/day, and sulfide, methane, and carbon dioxide were the main reaction products. Temporary inhibition of MT degradation occurred after MT peak loads and in the presence of dimethyl disulfide (DMDS), which is the autooxidation product of MT. From day 252 onwards, methanol was dosed to the reactor as co-substrate at a loading rate of 3-6 mmol/L/day to stimulate growth of methylotrophic methanogens. Methanol was completely degraded and also a complete MT degradation was achieved till a volumetric loading rate of 13 mmol MT/L/day (0.77 mmol MT/gVSS/day). However, from day 354 till the end of the experimental run (day 365), acetate was formed and MT was not completely degraded anymore, indicating that methanol-degrading homoacetogenic bacteria had partially outcompeted the methanogenic MT-degrading archea. The archeal community in the reactor sludge was analyzed by DGGE and sequencing of 16S rRNA genes. The methanogenic archea responsible for the degradation of MT in the reactor were related to Methanolobus oregonensis. A pure culture, named strain SODA, was obtained by serial dilutions in medium containing both trimethyl amine and dimethyl sulfide (DMS). Strain SODA degraded MT, DMS, trimethyl amine, and methanol. Flow sheet simulations revealed that for sufficient MT removal from liquefied petroleum gas, the extraction and biological degradation process should be operated above pH 9.  相似文献   

3.
The feasibility of anaerobic treatment of wastewater containing methanethiol (MT), an extremely volatile and malodorous sulfur compound, was investigated in lab-scale bioreactors. Inoculum biomass originating from full-scale anaerobic wastewater treatment facilities was used. Several sludges, tested for their ability to degrade MT, revealed the presence of organisms capable of metabolizing MT as their sole source of energy. Furthermore, batch tests were executed to gain a better understanding of the inhibition potential of MT. It was found that increasing MT concentrations affected acetotrophic organisms more dramatically than methylotrophic organisms. Continuous reactor experiments, using two lab-scale upflow anaerobic sludge bed (UASB) reactors (R1 and R2), aimed to determine the maximal MT load and the effect of elevated sulfide concentrations on MT conversion. Both reactors were operated at a hydraulic retention time (HRT) of about 7 hours, a temperature of 30 degrees C, and a pH of between 7.3 and 7.6. At the highest influent MT concentration applied, 14 mM in R1, corresponding to a volumetric loading rate of about 50 mM MT per day, 87% of the organic sulfur was recovered as hydrogen sulfide (12.2 mM) and the remainder as volatile organic sulfur compounds (VOSCs). Upon decreasing the HRT to 3.5 to 4.0 h at a constant MT loading rate, the sulfide concentration in the reactor decreased to 8 mM and MT conversion efficiency increased to values near 100%. MT conversion was apparently inhibited by the high sulfide concentrations in the reactor. The specific MT degradation rate, as determined after 120 days of operation in R1, was 2.83 +/- 0.27 mmol MT g VSS(-1) day(-1). During biological desulfurization of liquid hydrocarbon phases, such as with liquefied petroleum gas (LPG), the combined removal of hydrogen sulfide and MT is desired. In R2, the simultaneous addition of sodium sulfide and MT was therefore studied and the effect of elevated sulfide concentrations was investigated. The addition of sodium sulfide resulted in enhanced disintegration of sludge granules, causing significant washout of biomass. Additional acetate, added to stimulate growth of methanogenic bacteria to promote granulation, was hardly converted at the termination of the experimental period.  相似文献   

4.
The methane formation rate of 300 m(3) of sludge from a full scale biowaste reactor, that was stored without feeding for six weeks during a maintenance period, was about 60% of the methanogenic activity before maintenance. The 300 m(3) sludge was then pumped back into the biowaste reactor. On the third day, after refilling of the stored biowaste suspension, anaerobic conditions were obtained and feeding was started by addition of 36.1 m(3) of fresh biowaste suspension (=11.3 tons biowaste). The pH dropped from originally pH 7.7 to pH 7.3 and later on to pH 6.8, which was considered the minimum allowed pH for methanogenesis to recover. Maximum concentrations of acetate (1.78 gl(-1)), n-butyrate (0.57 gl(-1)) and n-valerate (0.44 gl(-1)) accumulated during the following days with feeding of 11.8 tons on day 5 and twice 6.5 tons on days 7 and 9, respectively. Thereafter, acetate, n-butyrate and n-valerate were degraded completely, whereas the concentration of propionate was still increasing. Propionic acid was the dominant fatty acid during the restart period and reached its maximum concentration of 6.2 gl(-1) 17 days after start of feeding. This high level of propionate was degraded completely in about 5 days with maximum degradation rates of 2.14 gl(-1)d(-1), and the pH of the anaerobic sludge increased from 7.1 to 7.4. During restart, the methane content of the biogas increased successively to 65%. Samples that were taken at different time intervals during the restart phase of the methane reactor showed different fatty acid degradation capabilities. After 10 days, when acetate and n-butyrate still accumulated in the methane reactor the maximum acetate degradation rate was 1.52 gl(-1)d(-1) and the n-butyrate degradation rate was 0.51 gl(-1)d(-1). Oxidation of n-valerate caused an increase of propionate, which was degraded after a lag phase of 6 days with a maximum rate of 0.6 gl(-1)d(-1). In the samples taken after 16 and 23 days, the propionate degradation rate increased to 1.42 gl(-1)d(-1) and 1.55 gl(-1)d(-1), respectively, and the lag phase for propionate degradation was reduced or had disappeared completely. The maximum propionate degradation rate was measured in the methane reactor in the fourth week after restart. The synthrophic propionate oxidizing bacteria were apparently the most suffering bacteria during sludge storage. If the propionate oxidizing bacteria could be kept active and the propionate degrading activity of the biowaste suspension of 6.16 gl(-1)d(-1) before the maintenance period could be maintained, then accumulation of 6.2 gl(-1) propionate in the methane reactor after restart could be avoided and full activity reached even earlier.  相似文献   

5.
Lon protease from Escherichia coli degraded lambda N protein in a reaction mixture consisting of the two homogeneous proteins, ATP, and MgCl2 in 50 mM Tris, Ph 8.0. Genetic and biochemical data had previously indicated that N protein is a substrate for Lon protease in vivo (Gottesman, S., Gottesman, M., Shaw, J. E., and Pearson, M. L. (1981) Cell 24, 225-233). Under conditions used for N protein degradation, several lambda and E. coli proteins, including native proteins, oxidatively modified proteins, and cloned fragments of native proteins, were not degraded by Lon protease. Degradation of N protein occurred with catalytic amounts of Lon protease and required the presence of ATP or an analog of ATP. This is the first demonstration of the selective degradation of a physiological substrate by Lon protease in vitro. The turnover number for N protein degradation was approximately 60 +/- 10 min-1 at pH 8.0 in 50 mM Tris/HCl, 25 mM MgCl2 and 4 mM ATP. By comparison the turnover number for oxidized insulin B chain was 20 min-1 under these conditions. Kinetic studies suggest that N protein (S0.5 = 13 +/- 5 microM) is intermediate between oxidized insulin B chain (S0.5 = 160 +/- 10 microM) and methylated casein (S0.5 = 2.5 +/- 1 microM) in affinity for Lon protease. N protein was extensively degraded by Lon protease with an average of approximately six bonds cleaved per molecule. In N protein, as well as in oxidized insulin B chain and glucagon, Lon protease preferentially cut at bonds at which the carboxy group was contributed by an amino acid with an aliphatic side chain (leucine or alanine). However, not all such bonds of the substrates were cleaved, indicating that sequence or conformational determinants beyond the cleavage site affect the ability of Lon protease to degrade a protein.  相似文献   

6.
The degradability of excess activated sludge from a wastewater treatment plant was studied. The objective was establishing the degree of degradation using either air or pure oxygen at different temperatures. Sludge treated with pure oxygen was degraded at temperatures from 22 degrees C to 50 degrees C while samples treated with air were degraded between 32 degrees C and 65 degrees C. Using air, sludge is efficiently degraded at 37 degrees C and at 50-55 degrees C. With oxygen, sludge was most effectively degraded at 38 degrees C or at 25-30 degrees C. Two-stage anaerobic-aerobic processes were studied. The first anaerobic stage was always operated for 5 days HRT, and the second stage involved aeration with pure oxygen and an HRT between 5 and 10 days. Under these conditions, there is 53.5% VSS removal and 55.4% COD degradation at 15 days HRT - 5 days anaerobic, 10 days aerobic. Sludge digested with pure oxygen at 25 degrees C in a batch reactor converted 48% of sludge total Kjeldahl nitrogen to nitrate. Addition of an aerobic stage with pure oxygen aeration to the anaerobic digestion enhances ammonium nitrogen removal. In a two-stage anaerobic-aerobic sludge digestion process within 8 days HRT of the aerobic stage, the removal of ammonium nitrogen was 85%.  相似文献   

7.
A Pseudomonas sp. strain NGKI (NCIM 5120) capable of degrading naphthalene was immobilized in polyurethane foam. The naphthalene-degrading activity of the freely suspended cells was compared with that of immobilized cells in batches in shaken culture and in a continuous culture system in a packed-bed reactor. Increasing concentrations of naphthalene were better tolerated and more quickly degraded by immobilized cell cultures than by free cells. An initial naphthalene concentration of 25 mM was completely degraded by freely suspended cells (4 x 10(10) cfu ml(-1)) and polyurethane-foam-immobilized cells (0.8-1 x 10(12) cfu g(-1) foam cubes) after 4 days and 2 days of incubation, respectively. Free cells degraded a maximum of 30 mM naphthalene after 4 days of incubation with 50 mM naphthalene, and no further degradation was observed even after 15 days of incubation, whereas foam-immobilized cells brought about the complete degradation of 50 mM initial naphthalene after 6 days of incubation. Furthermore, with 25 mM naphthalene, the polyurethane-foam-immobilized cells were re-used 45 times over a period of 90 days without losing naphthalene-degrading activity. By contrast, with the same amount of naphthalene, alginate-, agar-, and polyacrylamide-entrapped cells could be reused for 18, 12, and 23 times over a period of 44, 28, and 50 days, respectively. During continuous degradation in a packed-bed reactor, foam-immobilized cells degraded 80 mM naphthalene at a rate of 150 ml(-1) h(-1). With the same flow rate and 40 mM naphthalene, this system operated efficiently and continuously for about 120 days, whereas the packed-bed reactor with alginate-, agar-, and polyacrylamide-entrapped cells could be operated only for 45, 40, and 60 days respectively. Thus, more efficient degradation of naphthalene could be achieved by immobilizing cells of Pseudomonas sp. strain NGK1 in polyurethane foam, rather than in the other matrices tested.  相似文献   

8.
除草剂二甲戊灵的真菌降解及其特性研究   总被引:10,自引:0,他引:10  
富集分离了除草剂二甲戊灵降解真菌,并研究了其降解特性,结果表明,真菌可以降解二甲戊灵,利用富集培养的方法从环境中分离到16株能降解二甲戊灵的真菌。其中10株真菌5d内对100mg·L^-1二甲戊灵的降解率大于60%,以其中3株生理耐受能力强、降解能力高的真菌为例,研究了外加碳源浓度、初始pH值、二甲戊灵浓度和培养温度对真菌生长量和降解能力的影响,此3株真菌经鉴定分别属于土生曲霉组(Aspergillus terreus)、长梗串孢霉属(Monilochaetes)和烟色曲霉组(Aspergillus furnigatus),在外加碳源浓度为0.5%~1.0%的范围内,真菌生长量和降解率达到最大;在中性培养液中,3株真菌的生长量大,降解能力强;在浓度为100mg·L^-1时降解率和生长量都比较大,而绝对去除量随二甲戊灵浓度的提高而增加,在500mg·L^-1时达到最大;真菌的生长和降解需要适宜的温度,20~30℃培养时,降解率和生长量最大,可为农药污染治理及生产污水处理提供理论依据。  相似文献   

9.
Although several microorganisms that produce and degrade methanethiol (MT) and dimethyl sulfide (DMS) have been isolated from various habitats, little is known about the numbers of these microorganisms in situ. This study reports on the identification and quantification of microorganisms involved in the cycling of MT and DMS in freshwater sediments. Sediment incubation studies revealed that the formation of MT and DMS is well balanced with their degradation. MT formation depends on the concentrations of both sulfide and methyl group-donating compounds. A most-probable number (MPN) dilution series with syringate as the growth substrate showed that methylation of sulfide with methyl groups derived from syringate is a commonly occurring process in situ. MT appeared to be primarily degraded by obligately methylotrophic methanogens, which were found in the highest positive dilutions on DMS and mixed substrates (methanol, trimethylamine [TMA], and DMS). Amplified ribosomal DNA restriction analysis (ARDRA) and 16S rRNA gene sequence analysis of the total DNA isolated from the sediments and of the DNA isolated from the highest positive dilutions of the MPN series (mixed substrates) revealed that the methanogens that are responsible for the degradation of MT, DMS, methanol, and TMA in situ are all phylogenetically closely related to Methanomethylovorans hollandica. This was confirmed by sequence analysis of the product obtained from a nested PCR developed for the selective amplification of the 16S rRNA gene from M. hollandica. The data from sediment incubation experiments, MPN series, and molecular-genetics detection correlated well and provide convincing evidence for the suggested mechanisms for MT and DMS cycling and the common presence of the DMS-degrading methanogen M. hollandica in freshwater sediments.  相似文献   

10.
We report in this study that under certain conditions formaldehyde interacts with DNA and makes it more efficient for hybridization on nitrocellulose filters. Hybridization signals of formaldehyde-treated DNA are stronger (up to 10 fold) as compared with that of the heat- or alkali-denatured DNA. Various parameters of the DNA-formaldehyde reaction are optimized as follows: (a) 6 x SSC, 10% formaldehyde, 60 degrees C, 20-30 min, reaction volume 10-200 microliters or (b) 6 x SSC, 5% formaldehyde, 98 degrees C, 15 min, reaction volume 10-200 microliters. Treatment of agarose gels after electrophoresis with formaldehyde improved both the transfer of DNA and the efficiency of hybridization. The following conditions are recommended for gel treatment: denaturation in 0.3 N NaOH, 1 M NaCl followed by neutralization with 0.5 M phosphate buffer, pH 7.0, containing 10% formaldehyde at 60 degrees C for 20 min.  相似文献   

11.
Sulfate reduction outcompeted methanogenesis at 65 degrees C and pH 7.5 in methanol and sulfate-fed expanded granular sludge bed reactors operated at hydraulic retention times (HRT) of 14 and 3.5 h, both under methanol-limiting and methanol-overloading conditions. After 100 and 50 days for the reactors operated at 14 and 3.5 h, respectively, sulfide production accounted for 80% of the methanol-COD consumed by the sludge. The specific methanogenic activity on methanol of the sludge from a reactor operated at HRTs of down to 3.5 h for a period of 4 months gradually decreased from 0. 83 gCOD. gVSS(-1). day(-1) at the start to a value of less than 0.05 gCOD. gVSS(-1). day(-1), showing that the relative number of methanogens decreased and eventually became very low. By contrast, the increase of the specific sulfidogenic activity of sludge from 0. 22 gCOD. gVSS(-1). day(-1) to a final value of 1.05 gCOD. gVSS(-1). day(-1) showed that sulfate reducing bacteria were enriched. Methanol degradation by a methanogenic culture obtained from a reactor by serial dilution of the sludge was inhibited in the presence of vancomycin, indicating that methanogenesis directly from methanol was not important. H(2)/CO(2) and formate, but not acetate, were degraded to methane in the presence of vancomycin. These results indicated that methanol degradation to methane occurs via the intermediates H(2)/CO(2) and formate. The high and low specific methanogenic activity of sludge on H(2)/CO(2) and formate, respectively, indicated that the former substrate probably acts as the main electron donor for the methanogens during methanol degradation. As sulfate reduction in the sludge was also strongly supported by hydrogen, competition between sulfate reducing bacteria and methanogens in the sludge seemed to be mainly for this substrate. Sulfate elimination rates of up to 15 gSO(4)(2-)/L per day were achieved in the reactors. Biomass retention limited the sulfate elimination rate.  相似文献   

12.
A newly isolated methanogen, strain DMS1(T), is the first obligately anaerobic archaeon which was directly enriched and isolated from a freshwater sediment in defined minimal medium containing dimethyl sulfide (DMS) as the sole carbon and energy source. The use of a chemostat with a continuous DMS-containing gas stream as a method of enrichment, followed by cultivation in deep agar tubes, resulted in a pure culture. Since the only substrates utilized by strain DMS1(T) are methanol, methylamines, methanethiol (MT), and DMS, this organism is considered an obligately methylotrophic methanogen like most other DMS-degrading methanogens. Strain DMS1(T) differs from all other DMS-degrading methanogens, since it was isolated from a freshwater pond and requires NaCl concentrations (0 to 0.04 M) typical of the NaCl concentrations required by freshwater microorganisms for growth. DMS was degraded effectively only in a chemostat culture in the presence of low hydrogen sulfide and MT concentrations. Addition of MT or sulfide to the chemostat significantly decreased degradation of DMS. Transient accumulation of DMS in MT-amended cultures indicated that transfer of the first methyl group during DMS degradation is a reversible process. On the basis of its low level of homology with the most closely related methanogen, Methanococcoides burtonii (94.5%), its position on the phylogenetic tree, its morphology (which is different from that of members of the genera Methanolobus, Methanococcoides, and Methanohalophilus), and its salt tolerance and optimum (which are characteristic of freshwater bacteria), we propose that strain DMS1(T) is a representative of a novel genus. This isolate was named Methanomethylovorans hollandica. Analysis of DMS-amended sediment slurries with a fluorescence microscope revealed the presence of methanogens which were morphologically identical to M. hollandica, as described in this study. Considering its physiological properties, M. hollandica DMS1(T) is probably responsible for degradation of MT and DMS in freshwater sediments in situ. Due to the reversibility of the DMS conversion, methanogens like strain DMS1(T) can also be involved in the formation of DMS through methylation of MT. This phenomenon, which previously has been shown to occur in sediment slurries of freshwater origin, might affect the steady-state concentrations and, consequently, the total flux of DMS and MT in these systems.  相似文献   

13.
Although several microorganisms that produce and degrade methanethiol (MT) and dimethyl sulfide (DMS) have been isolated from various habitats, little is known about the numbers of these microorganisms in situ. This study reports on the identification and quantification of microorganisms involved in the cycling of MT and DMS in freshwater sediments. Sediment incubation studies revealed that the formation of MT and DMS is well balanced with their degradation. MT formation depends on the concentrations of both sulfide and methyl group-donating compounds. A most-probable number (MPN) dilution series with syringate as the growth substrate showed that methylation of sulfide with methyl groups derived from syringate is a commonly occurring process in situ. MT appeared to be primarily degraded by obligately methylotrophic methanogens, which were found in the highest positive dilutions on DMS and mixed substrates (methanol, trimethylamine [TMA], and DMS). Amplified ribosomal DNA restriction analysis (ARDRA) and 16S rRNA gene sequence analysis of the total DNA isolated from the sediments and of the DNA isolated from the highest positive dilutions of the MPN series (mixed substrates) revealed that the methanogens that are responsible for the degradation of MT, DMS, methanol, and TMA in situ are all phylogenetically closely related to Methanomethylovorans hollandica. This was confirmed by sequence analysis of the product obtained from a nested PCR developed for the selective amplification of the 16S rRNA gene from M. hollandica. The data from sediment incubation experiments, MPN series, and molecular-genetics detection correlated well and provide convincing evidence for the suggested mechanisms for MT and DMS cycling and the common presence of the DMS-degrading methanogen M. hollandica in freshwater sediments.  相似文献   

14.
The anaerobic degradation of terephthalate as sole substrate was studied in three anaerobic upflow reactors. Initially, the reactors were operated as upflow anaerobic sludge bed (UASB) reactors and seeded with suspended methanogenic biomass obtained from a full-scale down-flow fixed film reactor, treating wastewater generated during production of purified terephthalic acid. The reactors were operated at 30, 37, and 55 degrees C. The terephthalate removal capacities remained low in all three reactors (<4 mmolxL-1xday-1, or 1 g of chemical oxygen demand (COD)xL-1xday-1) due to limitations in biomass retention. Batch experiments with biomass from the UASB reactors revealed that, within the mesophilic temperature range, optimal terephthalate degradation is obtained at 37 degrees C. No thermophilic terephthalate-degrading culture could be obtained in either continuous or batch cultures. To enhance biomass retention, the reactors were modified to anaerobic hybrid reactors by introduction of two types of reticulated polyurethane (PUR) foam particles. The hybrid reactors were operated at 37 degrees C and seeded with a mixture of biomass from the UASB reactors operated at 30 and 37 degrees C. After a lag period of approximately 80 days, the terephthalate conversion capacity of the hybrid reactors increased exponentially at a specific rate of approximately 0.06 day-1, and high removal rates were obtained (40-70 mmolxL-1xday-1, or 10-17 g of CODxL-1xday-1) at hydraulic retention times between 5 and 8 h. These high removal capacities could be attributed to enhanced biomass retention by the development of biofilms on the PUR carrier material as well as the formation of granular biomass. Biomass balances over the hybrid reactors suggested that either bacterial decay or selective wash-out of the terephthalate fermenting biomass played an important role in the capacity limitations of the systems. The presented results suggest that terephthalate can be degraded at high volumetric rates if sufficiently long sludge ages can be maintained, and the reactor pH and temperature are close to their optima.  相似文献   

15.
A Pseudomonas sp. grew with nicotine optimally 3 g l(-1) and at 30 degrees C and pH 7. Nicotine was fully degraded within 10 h. The resting cells degraded nicotine in tobacco solid waste completely within 6 h in 0.02 m sodium phosphate buffer (pH 7) at maximally 56 mg nicotine h(-1) g dry cell(-1).  相似文献   

16.
Effect of pH on anoxic sulfide oxidizing reactor performance   总被引:3,自引:1,他引:3  
The effects of pH on the performance of anoxic sulfide oxidizing (ASO) reactor were evaluated. Performance was investigated under various operational conditions at influent pH range of 4-11. At the influent pH of 7-7.5 during loading tests and HRT tests, the sulfide oxidation was partial. In general, the amount of sulfate formed decreased with the increasing sulfide and nitrite loadings. The bacterial communities in ASO reactors were more sensitive to acidic pH compared with alkaline pH, as nitrite and sulfide removal rates dropped significantly when exposed to acidic pH 3. High dissolved bisulfide ions, nitrite and excess of sulfate (>300 mg/L) might have inhibited the sulfide oxidation under highly acidic and alkaline conditions in the ASO reactor. Based on sulfide and nitrite removal efficiencies, the ASO reactor can be operated in a wide range of pH, i.e. 5-11.  相似文献   

17.
S W King  V R Lum  T H Fife 《Biochemistry》1987,26(8):2294-2300
The carbamate ester N-(phenoxycarbonyl)-L-phenylalanine binds well to carboxypeptidase A in the manner of peptide substrates. The ester exhibits linear competitive inhibition toward carboxypeptidase A catalyzed hydrolysis of the amide hippuryl-L-phenylalanine (Ki = 1.0 X 10(-3) M at pH 7.5) and linear noncompetitive inhibition toward hydrolysis of the specific ester substrate O-hippuryl-L-beta-phenyllactate (Ki = 1.4 X 10(-3) M at pH 7.5). Linear inhibition shows that only one molecule of inhibitor is bound per active site at pH 7.5. The hydrolysis of the carbamate ester is not affected by the presence of 10(-8)-10(-9) M enzyme (the concentrations employed in inhibition experiments), but at an enzyme concentration of 3 X 10(-6) M catalysis can be detected. The value of kcat at 30 degrees C, mu = 0.5 M, and pH 7.45 is 0.25 s-1, and Km is 1.5 X 10(-3) M. The near identity of Km and Ki shows that Km is a dissociation constant. Substrate inhibition can be detected at pH less than 7 but not at pH values above 7, which suggests that a conformational change is occurring near that pH. The analogous carbonate ester O-(phenoxycarbonyl)-L-beta-phenyllactic acid is also a substrate for the enzyme. The Km is pH independent from pH 6.5 to 9 and has the value of 7.6 X 10(-5) M in that pH region. The rate constant kcat is pH independent from pH 8 to 10 at 30 degrees C (mu = 0.5 M) with a limiting value of 1.60 s-1. Modification of the carboxyl group of glutamic acid-270 to the methoxyamide strongly inhibits the hydrolysis of O-(phenoxycarbonyl)-L-beta-phenyllactic acid. Binding of beta-phenyllactate esters and phenylalanine amides must occur in different subsites, but the ratios of kcat and kcat/Km for the structural change from hippuryl to phenoxy in each series are closely similar, which suggests that the rate-determining steps are mechanistically similar.  相似文献   

18.
链霉菌Z94-2碱性脂肪酶产生条件及酶学性质   总被引:2,自引:0,他引:2  
在152株脂肪酶产生菌中,链霉菌Z94-2产脂肪酶活力为596u/mL,其最适培养基(g/L)为:糊精10、黄豆饼粉30、尿素10、K2HPO40.5、MgSO40.5、NaCl1和AEO90.5,产酶的最适条件为:初始pH9.5~10.0,在26℃培养48h。用PVA橄榄油乳化系统测定该酶的最适pH9.8,最适温度37℃,在pH8.6~10.2于5℃存放24h,酶活力不变。0.14mol/L的氯  相似文献   

19.
Branched alkanes including 2,6,10,14-tetramethylpentadecane (pristane) are more resistant to biological degradation than straight-chain alkanes especially under low-temperature conditions, such as 10 degrees C. Two bacterial strains, TMP2 and T12, that are capable of degrading pristane at 10 degrees C were isolated and characterized. Both strains grew optimally at 30 degrees C and were identified as Rhodococcus sp. based on the 16S rRNA gene sequences. Strain T12 degraded comparable amounts of pristane in a range of temperatures from 10 to 30 degrees C and strain TMP2 degraded pristane similarly at 10 and 20 degrees C but did not degrade it at 30 degrees C. These data suggest that the strains have adapted their pristane degradation system to moderately low-temperature conditions.  相似文献   

20.
An alkaliphilic Dietzia sp., strain GS-1, which degraded disodium terephthalate (DT), was isolated from soil. Strain GS-1 degraded 19.3 mM of DT in 168 h at pH 10. The maximum degradation velocity was 0.46 mM/h. The resting cells efficiently degraded 28.7 mM of DT in 51 h at 28 degrees C and pH 10. The degradation velocity was 0.41 mM/(h g-wet cell).  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