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.
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   

2.
The effect of dispersed n -dodecane or n -hexadecane on the air-to-aqueous phase overall volumetric oxygen transfer coefficient in a simulated (cell-free) stirred-tank fermentor is described. The oil volume fraction ranged from zero to 0.10; the ionic strength of the aqueous phases was varied from 0 to 0.45. The air-to-aqueous phase coefficients in both oil-free (KLa) and oil-bearing (KLa*) systems were evaluated from unsteady-state experiments using a membrane-covered probe to follow the aqueous phase dissolved oxygen tension. For all systems studied, KLa*/KLa was found to be independent of P/V and vs for all practical purposes. However, for a particular aqueous phase and at a given P/V and vs, the ratio KLa*KLa generally differed from unity. Depending on the combination of hydrocarbon type and volume fraction and the aqueous-phase ionic strength employed, the dispersed hydrocarbon may, in some cases, reduce the rate of oxygen transfer and in others enhance it relative to that of the corresponding oil-free gas–liquid dispersion. Enhancement of the air-to-aqueous transfer rate by such negative spreading coefficient hydrocarbons has not been reported previously.  相似文献   

3.
The activity of xenobiotic-degrading microorganisms is generally high in a biphasic aqueous/organic system. Therefore, the influence of interfacial area variation on kinetic parameters of Candida sp. growing on ethyl butyrate was evaluated. Interfacial areas of both aseptic and cultured biphasic systems were utilized. Substrate transport measurements in aseptic system (where the interface varied with the organic-phase fraction and agitator speed) showed that the substrate concentration in the aqueous phase was constant at different agitation speeds and decreased as the organic phase increased. Kinetic measurements of the cultured system showed that kinetic parameters vary as functions of their respective aseptic interfacial areas. Higher µmax and K i and lower K s values were obtained with larger interfacial areas. Measurements of the cultured system showed that the interfacial area increased as the biomass increased, and that about 50% of the biomass was attached to the interface as an interfacial biofilm at the end of the culture. Results suggest that the growth and selection of xenobiotic-degrading microorganisms in a biphasic aqueous/organic system should be evaluated mainly on the basis of the activity of adhering biomass (forming a biofilm) at the interfacial area rather than on substrate transport to the aqueous phase  相似文献   

4.
When cells of a type II methanotrophic bacterium (Methylocystis strain LR1) were starved of methane, both the Km(app) and the Vmax(app) for methane decreased. The specific affinity (aos) remained nearly constant. Therefore, the decreased Km(app) in starved cells was probably not an adjustment to better utilize low-methane concentrations.  相似文献   

5.
Methylobacterium sp. strain CRL-26 grown in a fermentor contained methane monooxygenase activity in soluble fractions. Soluble methane monooxygenase catalyzed the epoxidation/hydroxylation of a variety of hydrocarbons, including terminal alkenes, internal alkenes, substituted alkenes, branched-chain alkenes, alkanes (C1 to C8), substituted alkanes, branched-chain alkanes, carbon monoxide, ethers, and cyclic and aromatic compounds. The optimum pH and temperature for the epoxidation of propylene by soluble methane monooxygenase were found to be 7.0 and 40°C, respectively. Among various compounds tested, only NADH2 or NADPH2 could act as an electron donor. Formate and NAD+ (in the presence of formate dehydrogenase contained in the soluble fraction) or 2-butanol in the presence of NAD+ and secondary alcohol dehydrogenase generated the NADH2 required for the methane monooxygenase. Epoxidation of propylene catalyzed by methane monooxygenase was not inhibited by a range of potential inhibitors, including metal-chelating compounds and potassium cyanide. Sulfhydryl agents and acriflavin inhibited monooxygenase activity. Soluble methane monooxygenase was resolved into three components by ion-exchange chromatography. All three compounds are required for the epoxidation and hydroxylation reactions.  相似文献   

6.
Abstract

Oenococcus oeni CECT4730, which catalyses the asymmetric reduction of 2-octanone to (R)-2-octanol with high enantioselectivity, was further studied to exploit its potential for production of (R)-2-octanol in an aqueous/organic solvent biphasic system. Variables such as the volume ratio of aqueous to organic phase (Va/Vo), buffer pH, reaction temperature, shaking speed, co-substrates and the ratio of biocatalyst to substrate were examined with respect to the molar conversion, the initial reaction rate and the product enantiomeric excess (e.e.). Under the optimized conditions (Va/Vo=1:1 (v/v), buffer pH=8.0, reaction temperature=30°C, shaking speed=150 rev/min, ratio of glucose to biomass=5.4:l (w/w), ratio of biocatalyst to substrate=0.51:l (g/mol)), the highest space time yield of (R)-2-octanol, 24 mmol L?1 per h, and >98% product e.e. were obtained at a substrate concentration close to 1.0 mol L?1 after 24 h reduction.  相似文献   

7.
The synthesis of the protected dipeptide BocGlyPheOMe, has been modellised when working in an aqueousorganic biphasic system, with papain as a catalyst. The mathematical model takes into account that one of the substrates, PheOMe, has parallel hydrolysis reactions and that the reaction only takes place in the aqueous phase while the whole reaction system is biphasic. The reaction system has been modellised when working in batch as well as when working in fed-batch mode, achieving a good prediction of the product evolution for both working strategies. When working in fed-batch mode, the extension of the undesired parallel reactions has been diminished, the model has been used for a computer aided optimisation of the addition sequence of PheOMe. The results obtained led to a process operation strategy with a compromise between yield and productivity.List of Symbols [i] concentration of any component i - [i] aq concentration of i in the aqueous phase - [i] bi concentration of i in the biphasic system - [E] 0 initial concentration of enzyme - k e, kq first order kinetic constants - K A, KB equilibrium constants - r m maximum rate of reaction This worked was financed by the Interministerial Commission for Science and Technology (CICYT)from the Spanish Government under projects number BIO/88-370 and SAF92-0261-CO2-02.  相似文献   

8.
A hydrophilic ionic liquid, 1-butyl-3-methylimidazolium (L)-lactate ([bmim][lactate]), was successfully employed as a co-solvent to improve the biodehydrogenation of steroid 11β-hydroxyl medroxyprogesterone (HMP) to pregna-1,4-diene-6α-methyl-3,20-dione-11β,17α-dihydroxy (PDMDD) in an aqueous-organic biphasic system using Arthrobacter simplex UR106 cells. First, a suitable biphasic system was constituted, composed of toluene and Tris-HCl buffer (0.05 mol/L, pH 7.6) at a 0.4 (V org /V aq ) volumetric phase ratio. Investigation of biodehydrogenation with the addition of [bmim][lactate] to the aqueous phase was then performed. The results demonstrate that biodehydrogenation is more efficient in a biphasic system containing 2% [bmim][lactate] as a co-solvent compared to an aqueous-toluene two-phase system. Higher bioconversion (83.2% versus 56.4%) was observed after 16 h at a substrate concentration of 80 mg HMP in 8 mL of toluene.  相似文献   

9.
  For a mass-transfer-limited system, it was demonstrated that the volumetric ethene transfer coefficient (k l a) from gas to water could be enhanced by dispersing adequate amounts of a water-immiscible organic liquid, namely the perfluorocarbon FC40, in the aqueous phase. When 26% (v/v) FC40 was dispersed in a culture of Mycobacterium parafortuitum an enhancement of k l a, calculated on a total liquid volume basis, of 1.8 times was found. Steady-state experiments in the absence of microorganisms, however, showed a 1.2-fold enhancement of k l a at 18.5% (v/v) FC40. At all FC40 volume fractions tested, enhancement factors with cells were higher than enhancements without cells; apparently the microorganisms or their excretion products affected the interfacial areas or characteristic phase dimensions. Received: 4 December 1995 / Received revision: 7 June 1996 / Accepted: 10 June 1996  相似文献   

10.
We report for the first time the use of liquid-liquid counter-current chromatography (CCC) for the preparative scale fractionation of plasmid DNA. Almost complete fractionation of supercoiled and open circular plasmid DNA (6.9 kb) could be achieved using a phase system comprising 12.5% (w/w) PEG 600 and 18% (w/w) K2HPO4. Experiments were carried out on a Brunel J-type CCC machine (100 ml PTFE coil) at a mobile phase flow rate of 0.5 ml min– 1 and a rotational speed of 600 rpm. Compared to conventional HPLC techniques the capacity of CCC is not limited by the surface area of resin available for adsorption. Symbols: C b, Concentration of plasmid in lower phase (g ml–1); C t, Concentration of plasmid in upper phase (g ml–1); CV, Total volume of mobile phase present in the coil and connecting leads (ml); K, Equilibrium solute partition coefficient (K=C t/C b); OC, Open circular plasmid; SC, Supercoiled plasmid; S f, Percentage stationary phase retention (S f=V s/V c); t s, Time for phase separation (s); V b, Volume of bottom phase (ml); V c, Coil volume (ml); V m, Volume of mobile phase present in coil at equilibrium (ml); V r, Volume ratio of two phases (V r=V t/V b); V s, Volume stationary phase present in coil at equilibrium (ml); V t, Volume of top phase (ml); V tot, Total volume of phase system (ml).  相似文献   

11.
Whole-cell assays were used to measure the effect of dichloromethane and trichloroethylene on methane oxidation by Methylosinus trichosporium OB3b synthesizing the membrane-associated or particulate methane monooxygenase (pMMO). For M. trichosporium OB3b grown with 20 μM copper, no inhibition of methane oxidation was observed in the presence of either dichloromethane or trichloroethylene. If 20 mM formate was added to the reaction vials, however, methane oxidation rates increased and inhibition of methane oxidation was observed in the presence of dichloromethane and trichloroethylene. In the presence of formate, dichloromethane acted as a competitive inhibitor, while trichloroethylene acted as a noncompetitive inhibitor. The finding of noncompetitive inhibition by trichloroethylene was further examined by measuring the inhibition constants K iE and K iES. These constants suggest that trichloroethylene competes with methane at some sites, although it can bind to others if methane is already bound. Whole-cell oxygen uptake experiments for active and acetylene-treated cells also showed that provision of formate could stimulate both methane and trichloroethylene oxidation and that trichloroethylene did not affect formate dehydrogenase activity. The finding that different chlorinated hydrocarbons caused different inhibition patterns can be explained by either multiple substrate binding sites existing in pMMO or multiple forms of pMMO with different activities. The whole-cell analysis performed here cannot distinguish between these models, and further work should be done on obtaining active preparations of the purified pMMO. Received: 3 November 1998 / Accepted: 1 March 1999  相似文献   

12.
We investigated the block of KATP channels by glibenclamide in inside-out membrane patches of rat flexor digitorum brevis muscle. (1) We found that glibenclamide inhibited KATP channels with an apparent K i of 63 nm and a Hill coefficient of 0.85. The inhibition of KATP channels by glibenclamide was unaffected by internal Mg2+. (2) Glibenclamide altered all kinetic parameters measured; mean open time and burst length were reduced, whereas mean closed time was increased. (3) By making the assumption that binding of glibenclamide to the sulphonylurea receptor (SUR) leads to channel closure, we have used the relation between mean open time, glibenclamide concentration and K D to estimate binding and unbinding rate constants. We found an apparent rate constant for glibenclamide binding of 9.9 × 107 m −1 sec−1 and an unbinding rate of 6.26 sec−1. (4) Glibenclamide is a lipophilic molecule and is likely to act on sulfonylurea receptors from within the hydrophobic phase of the cell membrane. The glibenclamide concentration within this phase will be greater than that in the aqueous solution and we have taken this into account to estimate a true binding rate constant of 1.66 × 106 m −1 sec−1. Received: 7 July 1996/Revised: 4 October 1996  相似文献   

13.
Production of L-tryptophan from L-serine and indole catalyzed by Escherichia coli, immobilized in k-carrageenan gel beads, is technically feasible in the liquidimpelled loop reactor (LLR), using an organic solvent, e.g. n-dodecane.With L-serine in large excess intrinsic reaction kinetics is approximately first order with respect to indole, with a reaction constant of 8.5×10–5 m3 kg dw –1 s–1.The overall process kinetics is jointly controlled by intrinsic kinetics and by intraparticle mass transfer resistance, which can be quantified using an effectiveness factor.Mass transfer of indole from the organic to the aqueous phase and from the aqueous to the gel phase are relatively fast and thus have negligible influence in the overall process kinetics, under the operational conditions tested. However, they may become important if the process is intensified by increasing the cell concentration in the gel and/or the gel hold-up in the reactor.A simple model which includes indole mass balances over the aqueous and organic phases, mass transfer and reaction kinetics, with parameters experimentally determined in independent experiments, was successful in simulating L-tryptophan production in the LLR.List of Symbols a, b, c coefficients of the equilibrium curve for indole between organic and aqueous phases - A, B, C, D, E, F auxiliary variables used in liquid-liquid mass transfer studies - a x specific interfacial area referred to the volume of the aqueous phase (m–1) - A x interfacial area (m2) - a Y specific interfacial area referred to the volume of the organic phase (m–1) - A Y interfacial area (m2) - C b substrate concentration in the bulk of the aqueous phase (kg m–3) - C e substrate concentration in exit stream (kg m–3) - C E biocatalyst concentration referred to the aqueous phase (kg m–3) - C E s biocatalyst concentration referred to the volume of gel (kg m–3) - C s substrate concentration at the gel surface (kgm–3) - d, e, f coefficients of the equilibrium curve for indole between aqueous and organic phases - dp particle diameter (m) - K 2 kinetic constant (s–1) - K 1 kinetic constant K2/KM (kg–1 m3 s–1) - K M Michaälis-Menten constant (kgm–3) - K X mass transfer coefficient referred to the aqueous phase (ms–1) - K XaX volumetric mass transfer coefficient based on the volume of the aqueous phase (s–1) - k Y mass transfer coefficient referred to the organic phase (ms–1) - K YaY volumetric mass transfer coefficient based on the volume of the organic phase (s–1) - N X mass flux of indole from organic to aqueous Phase (kg m–2s–1) - N Y mass flux of indole from aqueous to organic phase (kg m–2s–1) - Q e volumetric flow rate in exit stream (m3s–1) - Q f volumetric flow rate in feed stream (m3s–1) - obs observed reaction rate (kg s–1 m–3) - intrinsic reaction rate (kg s–1 m–3) - Re Reynolds number - Sc Schmidt number - Sh Sherwood number - t time (s) - u superficial velocity (m s–1) - V max maximum reaction rate (kg s–1m–3) - V S volume of the support (m3) - V X volume of aqueous phase (m3) - V Y volume of the organic phase (m3) - X indole concentration in the aqueous phase (kgm–3) - Y indole concentration in the organic phase (kg m–3 Greek Letters overall effectiveness factor - e external effectiveness factor - i internal effectiveness factor - Thiele module A fellowship awarded to one of us (D.M.R.)by INICT is gratefuly acknowledged.  相似文献   

14.
 In order to enhance the productivity of lactic acid and reduce the end-product inhibition of fermentation, the partitioning and growth of four different strains of lactic acid bacteria in three different aqueous two-phase systems were studied. Polyethyleneglycol/ dextran, polyethyleneglycol/hydroxypropyl starch polymer (HPS), and a random copolymer of ethylene oxide and propylene oxide (EO-PO)/HPS were used as polymer systems. One strain each of Lactococcus lactis subsp. lactis and of Lactobacillus delbrueckii subsp. delbrueckii partitioned completely to the interface and bottom phase in two-phase systems with low polymer concentrations of EO-PO/HPS100 and EO-PO/ HPS200. The growth and production of lactic acid by two of three L. lactis strains in a two-phase system with 5.5% (w/w) EO-PO and 12.0% (w/w) HPS100 were reduced by less than 10% compared with a reference fermentation in a normal growth medium. The viability of L. lactis subsp. lactis ATCC 19435 was maintained for at least 50 h and with four top-phase replacements during extractive fermentation in the EO-PO/HPS100 system. Moreover, when cell density reached the stationary phase in the first extractive fermentation, the lactate production in this aqueous two-phase system was maintained. Received: 2 October 1995/Received revision: 16 January 1996/Accepted: 22 January 1996  相似文献   

15.
Whole-cell assays of methane and trichloroethylene (TCE) consumption have been performed on Methylosinus trichosporium OB3b expressing particulate methane monooxygenase (pMMO). From these assays it is apparent that varying the growth concentration of copper causes a change in the kinetics of methane and TCE degradation. For M. trichosporium OB3b, increasing the copper growth concentration from 2.5 to 20 μM caused the maximal degradation rate of methane (Vmax) to decrease from 300 to 82 nmol of methane/min/mg of protein. The methane concentration at half the maximal degradation rate (Ks) also decreased from 62 to 8.3 μM. The pseudo-first-order rate constant for methane, Vmax/Ks, doubled from 4.9 × 10−3 to 9.9 × 10−3 liters/min/mg of protein, however, as the growth concentration of copper increased from 2.5 to 20 μM. TCE degradation by M. trichosporium OB3b was also examined with varying copper and formate concentrations. M. trichosporium OB3b grown with 2.5 μM copper was unable to degrade TCE in both the absence and presence of an exogenous source of reducing equivalents in the form of formate. Cells grown with 20 μM copper, however, were able to degrade TCE regardless of whether formate was provided. Without formate the Vmax for TCE was 2.5 nmol/min/mg of protein, while providing formate increased the Vmax to 4.1 nmol/min/mg of protein. The affinity for TCE also increased with increasing copper, as seen by a change in Ks from 36 to 7.9 μM. Vmax/Ks for TCE degradation by pMMO also increased from 6.9 × 10−5 to 5.2 × 10−4 liters/min/mg of protein with the addition of formate. From these whole-cell studies it is apparent that the amount of copper available is critical in determining the oxidation of substrates in methanotrophs that are expressing only pMMO.  相似文献   

16.

Biotransformation is a green and useful tool for sustainable and selective chemical synthesis. However, it often suffers from the toxicity and inhibition from organic substrates or products. Here, we established a hollow fiber membrane bioreactor (HFMB)-based aqueous/organic biphasic system, for the first time, to enhance the productivity of a cascade biotransformation with strong substrate toxicity and inhibition. The enantioselective trans-dihydroxylation of styrene to (S)-1-phenyl-1,2-ethanediol, catalyzed by Escherichia coli (SSP1) coexpressing styrene monooxygenase and an epoxide hydrolase, was performed in HFMB with organic solvent in the shell side and aqueous cell suspension in the lumen side. Various organic solvents were investigated, and n-hexadecane was found as the best for the HFMB-based biphasic system. Comparing to other reported biphasic systems assisted by HFMB, our system not only shield much of the substrate toxicity but also deflate the product recovery burden in downstream processing as the majority of styrene stayed in organic phase while the diol product mostly remained in the aqueous phase. The established HFMB-based biphasic system enhanced the production titer to 143 mM, being 16-fold higher than the aqueous system and 1.6-fold higher than the traditional dispersive partitioning biphase system. Furthermore, the combination of biphasic system with HFMB prevents the foaming and emulsification, thus reducing the burden in downstream purification. HFMB-based biphasic system could serve as a suitable platform for enhancing the productivity of single-step or cascade biotransformation with toxic substrates to produce useful and valuable chemicals.

  相似文献   

17.
Summary The kinetics of methane uptake by Methylococcus capsulatus (Bath) and its inhibition by ammonia were studied by stopped-flow membrane-inlet mass spectrometry. Measurements were done on suspensions of cells grown in high- and low-copper media. With both types of cells the kinetics of methane uptake are hyperbolic when oxygen is in excess. The apparent K m and K max for methane uptake are both higher in low-copper cells than in high-copper cells. Ammonia is a simple competitive inhibitor of methane uptake in high-copper cells when the oxygen concentration is above a few M. The findings agree with the assumption that ammonia is a week alternative substrate for particulate methane monooxygenase. In low-copper cells the effect of ammonia is complicated and cannot be explained in terms of current assumptions on the mechanism of soluble methane monooxygenase. Our data indicate that ammonia inhibition is likely to be a more serious problem in connection with cultivation in low-copper medium than in high-copper medium. Offprint requests to: H. N. Carlsen  相似文献   

18.
W/O微乳液中糖化酶的研究   总被引:2,自引:0,他引:2  
在十六烷基三甲基溴化铵/正戊醇/异辛烷/水体系中,选取W/O型微乳液作为糖化酶的微环境,研究了其催化活性.并与其在水溶液中进行了比较.结果表明,在微乳液中,糖化酶催化淀粉水解反应的最佳反应温度和pH比在水溶液中低,反应的最大速度Vm和米氏常数Km比其在水溶液中分别提高了近6倍和3倍.  相似文献   

19.
Poly(3-hydroxybutyrate) depolymerase was purified to homogeneity from the culture filtrate of Paecilomyces lilacinus D218 by column chromatography on CM-Toyopearl 650M and hydroxylapatite. The molecular weight of the enzyme was estimated to be 48,000 by SDS-PAGE. Maximal activity was observed near pH 7.0 and 45°C. The K m and V max values for PHB were 0.13 (mg/ml) and 3750 (U/mg protein), respectively. The enzyme hydrolyzed PHB and p-nitrophenyl fatty acids but not polycaprolactone and triglycerides. Received: 29 August 1996 / Accepted: 30 September 1996  相似文献   

20.
Soluble methane monooxygenase (sMMO) maximization studies were carried out as part of a larger effort directed towards the development and optimization of an aqueous phase, multistage, membrane bioreactor system for treatment of polluted groundwater. A modified version of the naphthalene oxidation assay was utilized to determine the effects of methane:oxygen ratio, nutrient supply, and supplementary carbon sources on maximizing and maintaining sMMO activity inMethylosinus trichosporium OB3b.Methylosinus trichosporium OB3b attained peak sMMO activity (275–300 nmol of naphthol formed h–1 mg of protein–1 at 25°C) in early stationary growth phase when grown in nitrate mineral salts (NMS) medium. With the onset of methane limitation however, sMMO activity rapidly declined. It was possible to define a simplified nitrate mineral salts (NMS) medium, containing nitrate, phosphate and a source of iron and magnesium, which allowed reasonably high growth rates (max 0.08 h–1) and growth yields (0.4–0.5 g cells/g CH4) and near maximal activities of sMMO. In long term batch culture incubations sMMO activity reached a stable plateau at approximately 45–50% of the initial peak level and this was maintained over several weeks. The addition of d-biotin, pyridoxine, and vitamin B12 (cyanocobalamin) increased the activity level of sMMO in actively growing methanotrophs by 25–75%. The addition of these growth factors to the simplified NMS medium was found to increase the plateau sMMO level in long term batch cultures up to 70% of the original peak activity.Abbreviations sMMO soluble methane monooxygenase - pMMO particulate methane monooxygenase - NMS nitrate mineral salts - TCE trichloroethene - NADH reduced nicotinamide adenine dinucleotide  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