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 influence of small doses of ethanol or dimethylsulphoxide (DMSO) on in vitro maturation (IVM) of bovine cumulus-oocyte complexes (COC) was examined, either after spontanous maturation or after inhibition of meiosis with 6-dimethylaminopurine (6-DMAP) or 3-isobutyl-1-methylxanthine (IBMX). Subsequent to IVM for 23 hr in semidefined serum-free Earle's TCM199 medium, nuclear maturation was assessed cytogenetically, while the combined cytoplasmic and nuclear maturation was measured indirectly by the oocytes' ability to undergo fertilization and further embryonic development. Embryo development was followed until the blastocyst stages at day 9 after insemination. Neither spontanous nuclear maturation nor cleavage was compromised by IVM in 相似文献   

2.
Cytochrome p450 BM-3 (EC 1.14.14.1) catalyzes the hydroxylation and/or epoxidation of a broad range of substrates, including alkanes, alkenes, alcohols, fatty acids, amides, polyaromatic hydrocarbons, and heterocycles. For many of these notoriously water-insoluble compounds, p450 BM-3's K(m) values are in the millimolar range. Polar organic cosolvents are therefore added to increase substrate solubility and achieve high catalytic efficiency. Using p450 BM-3 as a catalyst for these important transformations requires that we improve its ability to tolerate the cosolvents. By directed evolution, we improved the activity of p450 BM-3 in the presence of dimethylsulfoxide (DMSO) and tetrahydrofuran (THF), achieving increases in specific activity up to 10-fold in 2% (v/v) THF and 6-fold in 25% (v/v) DMSO. The engineered p450 BM-3's are also significantly more resistant to acetone, acetonitrile, dimethylformamide, and ethanol as cosolvents in the reaction.  相似文献   

3.
The activity and stability of horseradish (Amoracia rusticana) peroxidase (HRP) free in solution and immobilised onto silica microparticles was studied in the presence of organic co-solvents.

The effect of several hydrophilic organic solvents, namely dimethyl sulfoxide, dimethylformamide, dioxan, acetonitrile and tetrahydrofuran, in the activity and stability of free HRP was studied. From the solvents tested, DMSO led to the highest activities and stabilities. After 2 h of incubation at 35°C, the remaining activity of the enzyme in the presence of 30% of each solvent was less than 30%, with exception of DMSO for which the enzyme remained fully active.

In order to increase stability, HRP was covalently immobilised onto silica microparticles. The half-life of the enzyme in buffer at 50°C increased from 2 to 52 h when the enzyme was immobilised. The stability of both free and immobilised HRP was also studied at 50°C in aqueous mixtures of 3.5, 20, 35 and 50% (v/v) DMSO. Free HRP stability was not affected by the presence of 3.5 and 20% DMSO, but higher contents lead to a more pronounced deactivation. Immobilised HRP stability increased with DMSO content up to 20%, decreasing for higher contents. The enzyme half-life increased more than 300% when changing from buffer to 20% DMSO.

The deactivation of free HRP was modelled using the simple exponential decay, and the deactivation of immobilised HRP was described by a two-step inactivation model.  相似文献   


4.
Abstract Intracellular acidification has been considered one of a number of mechanisms underlying the inhibition of growth and fermentation by ethanol in yeast. However, most of the studies on the effect of ethanol on yeast intracellular pH (pHi) were carried out by using unadapted cells to which ethanol was added. In this paper we show that the pHi of exponential cells of Saccharomyces cerevisiae IGC 3507 III grown in a medium with glucose and inhibitory concentrations of ethanol only decreased to values below those in unstressed cells (6.9) for concentrations equal to or above 7% (v/v). Only at these supracritical levels (7–10% (v/v)) was pH homeostasis in ethanol-adapted yeast affected. This is consistent with the significant increase of plasma membrane permeability and decrease of plasma membrane H+-ATPase in comparison with the corresponding values in unstressed cells. These deleterious effects were only observed with those high concentrations of toxin. These results indicate that intracellular acidification does not account for inhibition of yeast growth in the presence of ethanol. In fact, growth was inhibited by ethanol concentrations (3–6% (v/v)) that did not lead to the decrease of pHi. Furthermore, even for supracritical concentrations, close to the maximal that allowed growth (10% (v/v)), the dedrease of pHi was not important reaching, at the most, values of 6.5–6.6.  相似文献   

5.
Wheat straw consists of 48.57 ± 0.30% cellulose and 27.70 ± 0.12% hemicellulose on dry solid (DS) basis and has the potential to serve as a low cost feedstock for production of ethanol. Dilute acid pretreatment at varied temperature and enzymatic saccharification were evaluated for conversion of wheat straw cellulose and hemicellulose to monomeric sugars. The maximum yield of monomeric sugars from wheat straw (7.83%, w/v, DS) by dilute H2SO4 (0.75%, v/v) pretreatment and enzymatic saccharification (45 °C, pH 5.0, 72 h) using cellulase, β-glucosidase, xylanase and esterase was 565 ± 10 mg/g. Under this condition, no measurable quantities of furfural and hydroxymethyl furfural were produced. The yield of ethanol (per litre) from acid pretreated enzyme saccharified wheat straw (78.3 g) hydrolyzate by recombinant Escherichia coli strain FBR5 was 19 ± 1 g with a yield of 0.24 g/g DS. Detoxification of the acid and enzyme treated wheat straw hydrolyzate by overliming reduced the fermentation time from 118 to 39 h in the case of separate hydrolysis and fermentation (35 °C, pH 6.5), and increased the ethanol yield from 13 ± 2 to 17 ± 0 g/l and decreased the fermentation time from 136 to 112 h in the case of simultaneous saccharification and fermentation (35 °C, pH 6.0).  相似文献   

6.
Due to its high content of lactose and abundant availability, cheese whey powder (CWP) has received much attention for ethanol production in fermentation processes. However, lactose‐fermenting yeast strains including Kluyveromyces marxianus can only produce alcohol at a relatively low level, while the most commonly used distiller yeast strain Saccharomyces cerevisiae cannot ferment lactose since it lacks both β‐galactosidase and the lactose permease system. To combine the unique aspects of these two yeast strains, hybrids of K. marxianus TY‐22 and S. cerevisiae AY‐5 were constructed by protoplast fusion. The fusants were screened and characterized by DNA content, β‐galactosidase activity, ethanol tolerance, and ethanol productivity. Among the genetically stable fusants, the DNA content of strain R‐1 was 6.94%, close to the sum of the DNA contents of TY‐22 (3.99%) and AY‐5 (3.51%). The results obtained by random‐amplified polymorphic DNA analysis suggested that R‐1 was a fusant between AY‐5 and TY‐22. During the fermentation process with CWP, the hybrid strain R‐1 produced 3.8% v/v ethanol in 72 h, while the parental strain TY‐22 only produced 3.1% v/v ethanol in 84 h under the same conditions.  相似文献   

7.
In this work a new highly fluorescent N,N‐dimethyl benzylamine–palladium(II) yu complex was synthesized by the reaction of [Pd2{(C,N–C6H4CH2N(CH3)2}2(μ‐OAc)]2] with curcumin. The structure of the synthesized complex was characterized using Fourier transform infra‐red (FT‐IR) spectroscopy, 1H nuclear magnetic resonance spectroscopy, and elemental analysis. Fluorescence quantum yield (ΦF) values of the synthesized complex in dimethyl sulfoxide (DMSO), acetonitrile, ethanol, and methanol were 0.160, 0.104, 0.068, and 0.061, respectively. The fluorescence signal of the complex in the organic solvents was very sensitive to the water content of the organic solvent. The quenching effect of water was used to determine trace amounts of water in the heteroatom‐containing organic solvents (ethanol, methanol, acetonitrile) and redox‐active solvents (DMSO). The linear ranges for determination of water (v/v %) in ethanol, DMSO and acetonitrile were found to be 0.03–14.5, 0.08–13.8, and 0.07–18.8, respectively. Two linear ranges were found for determination of water (v/v %) in methanol (0.1–1.2 and 4.7–25.0). Detection limit (DL) values were calculated to be 0.001, 0.05, 0.004, and 0.01 (v/v %) in ethanol, methanol, acetonitrile, and DMSO, respectively. The proposed method overcomes the problems of the standard Karl Fischer method for determination of water in DMSO. In addition, it gave the best DL value for determination of water in ethanol compared with all published papers to date.  相似文献   

8.
Pepsin, a member of the aspartate protease family, exists in a partially unfolded state at alkaline pH where the N-terminal domain of pepsin has a flexible structure while the C-terminal domain has a highly folded structure. In this work, the conformational stability of porcine pepsin in an alkaline denatured (A(D)) state against acetonitrile and ethanol solvents was studied using a combination of electronic circular dichroism (ECD) and fluorescence techniques. The ECD results demonstrate that both ethanol and acetonitrile induce secondary structural changes in pepsin at A(D) state. However, the minimum concentration required to induce significant secondary structural changes in pepsin varies for ethanol (>30%, v/v) and acetonitrile (>60%, v/v) solvents. At maximum concentration used (90%, v/v), both solvents induce predominantly β-sheet conformation. Unlike acetonitrile, ethanol induces significant amount of non-native α-helical conformations at the intermediate concentrations (50-80%). The tryptophan fluorescence results demonstrate that both acetonitrile and ethanol induce substantial changes in the tertiary structure of pepsin in the A(D) state above certain concentrations. The current results have important implications in understanding the effect of co-solvents on the conformation of proteins in the "denatured state".  相似文献   

9.
Condensation of Cbz-Asp and PheOMe catalyzed by a neutral protease from Vibrio sp. T1800 (Vimelysin: VLN) was studied. VLN showed a relatively higher catalytic activity of condensation and an apparently larger yield after 3 h or 24 h, in comparison with thermolysin (TLN), especially at lower pH and temperatures.VLN showed higher solvent-tolerance than TLN. TLhe apparent highest yield (25%) was obtained in 30% DMSO by using VLN; under similar conditions, TLN gave only about a half of this value. The rate of the condensation reaction per mole of enzyme (v/[E](o)) in DMSO 50% at 37 degrees C and pH 6.5 was 0.16 s(-1) for VLN and 0.047 s(-1) for TLN. In 30% ethanol VLN showed more than three-fold peptide yield than TLN after 5 h reaction. (c) 1997 John Wiley & Sons, Inc. Biotechnol Bioeng 53: 387-390, 1997.  相似文献   

10.
The influence of three parameters on the degree of acquired tolerance to ethanol was studied using goldfish as the animal model. These factors were: a) the rate of ethanol administration; b) the repeated performance of the behavioral test, and c) the prior presence of tolerance to a low dose of ethanol. Tolerance, as indicated by the increase in the ethanol brain levels associated with the loss of righting reflex (overturn test), was observed in fish exposed to 0.2, 0.4, or 0.6% (w/v) ethanol for 1 to 24 hours independently of the rate of ethanol administration. Fish exposed to 0.8% (w/v) ethanol developed tolerance only when this ethanol concentration was reached slowly (circa 3 hours). The increase in functional demand by the repeated testing of the fish and the presence of tolerance to a low dose of ethanol (0.4%) did not affect the degree of the observed tolerance in fish left to equilibrate with a 0.8% ethanol dose. The data support the suggestion that motor functions controlled by the central nervous system can adapt to a high concentration of ethanol if an adequate amount of time is allowed for this compensation to occur or if this concentration is reached slowly. Furthermore, in goldfish the degree of acquired tolerance is directly proportional to the concentration of the ethanol during the time of exposure and independent of the previous experience in alcohol solution.  相似文献   

11.
Zhang L  Zhang M  Dong J  Guo J  Song Y  Cheung PC 《Biopolymers》2001,59(6):457-464
A water-insoluble polysaccharide (TM8) was isolated from sclerotium of Pleurotus tuber-regium by extraction with 0.5M NaOH aqueous solutions at 120 degrees C. Its chemical structure was confirmed by infrared, high performance liquid chromatography, gas chromatography, and (13)C NMR in dimethylsulfoxide (DMSO) to be composed of beta-(1 --> 3)-D-glucan backbone chain linked with a branched glucose, one out of every three glycosyl units being substituted at C6 position. The glucan TM8 in DMSO was fractionated by nonsolvent addition method into ten fractions, and the solution properties were studied by size exclusion chromatography combined with multiangle laser light scattering (SEC-MALLS) and viscometry in DMSO at 30 degrees C. The dependencies of intrinsic viscosity [eta] and radius of gyration [(s(2)(1/2)(z-2)] on weight-average molecular mass M(w) for this glucan were found to be [eta] = (9.24 +/- 0.2) x 10(-2)M(w)(0.51 +/- 0.02) (cm(3)g(-1)) and [(s(2)(1/2)(z-2)] = (3.67 +/- 0.3) x 10(-2)M(w)(0.56 +/- 0.02) (nm) in the range of M(w) from 1.07 x 10(4) to 77.4 x 10(4). Based on current theories for a wormlike chain, the conformational parameters of the glucan TM8 were found to be 408 (nm(-1)) for M(L), 3.1 (nm) for q, and 16.8 for C(infinity), suggesting that the polysaccharide exists as a dense random-coil chain in DMSO, due to branched structure.  相似文献   

12.
The low ethanol tolerance of thermophilic anaerobic bacteria (<2%, v/v) is a major obstacle for their industrial exploitation for ethanol production. The ethanol tolerance of the thermophilic anaerobic ethanol-producing strain Thermoanaerobacter A10 was studied during batch tests of xylose fermentation at a temperature range of 50-70 degrees C with exogenously added ethanol up to approximately 6.4% (v/v). At the optimum growth temperature of 70 degrees C, the strain was able to tolerate 4.7% (v/v) ethanol, and growth was completely inhibited at 5.6% (v/v). A higher ethanol tolerance was found at lower temperatures. At 60 degrees C, the strain was able to tolerate at least 5.1% (v/v) ethanol. A generalized form of Monod kinetic equation proposed by Levenspiel was used to describe the ethanol (product) inhibition. The model predicted quite well the experimental data for the temperature interval 50-70 degrees C, and the maximum specific growth rate and the toxic power (n), which describes the order of ethanol inhibition at each temperature, were estimated. The toxic power (n) was 1.33 at 70 degrees C, and corresponding critical inhibitory product concentration (P(crit)) above which no microbial growth occurs was determined to be 5.4% (v/v). An analysis of toxic power (n) and P(crit) showed that the optimum temperature for combined microbial growth and ethanol tolerance was 60 degrees C. At this temperature, the toxic power (n), and P(crit) were 0.50, and 6.5% (v/v) ethanol, respectively. From a practical point of view, the model may be applied to compare the ethanol inhibition (ethanol tolerance) on microbial growth of different thermophilic anaerobic bacterial strains.  相似文献   

13.
A new epoxide hydrolase with high enantioselectivity toward (R)-glycidyl phenyl ether (R-GPE) was partially purified from Bacillus megaterium strain ECU1001. The maximum activity of the isolated enzyme was observed at 30 degrees C and pH 6.5 in a buffer system with 5% (v/v) of DMSO as a cosolvent. The enzyme was quite stable at pH 7.5 and retained full activity after incubation at 40 degrees C for 6 h. Interestingly, when the cosolvent DMSO was replaced by an emulsifier (Tween-80, 0.5% w/v) as an alternative additive to help disperse the water-insoluble substrate, the apparent activity of the epoxide hydrolase significantly increased by about 1.8-fold, while the temperature optimum shifted from 30 to 40 degrees C and the half-life of the enzyme at 50 degrees C increased by 2.5 times. The enzymatic hydrolysis of rac-GPE was highly enantioselective, with an E-value (enantiomeric ratio) of 69.3 in the Tween-80 emulsion system, which is obviously higher than that (41.2) observed in the DMSO-containing system.  相似文献   

14.
The effect of solvents and solvent mixtures on the synthesis of myristic acid esters of different carbohydrates with an immobilized lipase from C. antarctica was investigated. The rate of myristyl glucose synthesized by the enzyme was increased from 3.7 to 20.2 micromol min(-1) g(-1) by changing the solvent from pure tert-butanol to a mixture of tert-butanol:pyridine (55:45 v/v), by increasing the temperature from 45 degrees C to 60 degrees C, and by optimizing the relative amounts of glucose, myristic acid, and the enzyme preparation. Addition of more than 2% DMSO to the tert-butanol:pyridine system resulted in a reduction of enzyme activity. Lowering the water content of the enzyme preparation below 0.85% (w/w) resulted in significant decreases in enzyme activity, while increasing the water content up to 2.17% (w/w) did not significantly affect the enzyme activity. The highest yields of myristyl glucose were obtained when an excess of unsolubilized glucose was present in the reaction system. In this case, all of the initially solubilized and a significant amount of the initially unsolubilized glucose was converted to the ester within 24 h of incubation, resulting in a myristyl glucose concentration of 34 mg/mL(-1). Myristic acid esters of fructose (22.3 micromol min(-1) g(-1)), alpha-D-methyl-glucopyranoside (26.9 micromol min(-1) g(-1)) and maltose (1.9 micromol min(-1) g(-1)) could also be prepared using the tert-butanol:pyridine solvent system. No synthesis activity was observed with maltotriose, cellobiose, sucrose, and lactose as substrate.  相似文献   

15.
Summary Continuous cultivation of Rhodotorula sp. Y-38 was carried out on ethanol, acetic acid or acetaldehyde. At a feed concentration of 1.0 % (w/v) ethanol, the cell yield of 64 g/100 g ethanol and crude protein of 52 g/100 g biomass were obtained at D=0.5 h-1. The respective value of the content of amino acids and nucleic acids was 42.6 and 9.4 g/100 g biomass. At 2.0 % (w/v) acetic acid, cell yield was found to be 50 g/100 g acetic acid at D=0.4 h-1. The optimum dilution rate ranged between 0.3 and 0.4 h-1. At 0.05 % (w/v) acetaldehyde, the maximum cell yield was obtained at D=0.14 h-1.  相似文献   

16.
Aims:  To study the effect of ethanol on Oenococcus oeni activity at the single cell level.
Methods and Results:  The active extrusion of the fluorescent probe carboxy fluorescein (cF) was used to assess the metabolic activity of ethanol-stressed O. oeni cells. Subsequent flow cytometric analysis revealed that O. oeni cells extrude the accumulated cF upon energizing with l -malic acid. However, O. oeni cells exposed to 12% (v/v) ethanol for 1 h showed a decreased capacity for active extrusion of cF. Moreover, two subpopulations could be distinguished, one of which being able to extrude cF and the other one remaining cF fluorescent. Growing cells in the presence of 8% (v/v) ethanol resulted in robust cells that maintained the capacity to actively extrude cF after being exposed to 12% (v/v) ethanol, which in turn correlated with the high levels of ATP observed in these ethanol stressed, malolactic fermentation (MLF) performing cells.
Conclusion:  From our results, it becomes evident that active extrusion of cF can be used to assess malolactic activity in O. oeni .
Significance and Impact of the Study:  The present study provides information for the development of a rapid method to assess the malolactic activity of individual O. oeni cells performing MLF during wine production.  相似文献   

17.
The thermotolerant, ethanol-producing yeast strain Kluyveromyces marxianus IMB3 was grown at 45°C on media containing 2, 4 and 6 % (w/v) pulverized barley straw and supplemented with 2% (v/v) cellulase. Maximum ethanol concentrations produced were 2, 3 and 3.6g/l, respectively. When the pulverized straw was replaced by NaOH pretreated straw (at 2, 4 and 6% (w/v); based on original untreated straw), ethanol concentrations increased to maxima of 3.9, 8, and 12g/l, respectively. The ethanol yields amount to 20g ethanol from 100g of straw.  相似文献   

18.
Simultaneous saccharification and fermentation (SSF) process for ethanol production from various lignocellulosic woody (poplar and eucalyptus) and herbaceous (Sorghum sp. bagasse, wheat straw and Brassica carinata residue) materials has been assayed using the thermotolerant yeast strain Kluyveromyces marxianus CECT 10875. Biomass samples were previously treated in a steam explosion pilot plant to provide pretreated biomass with increased cellulose content relative to untreated materials and to enhance cellulase accessibility. SSF experiments were performed in laboratory conditions at 42 °C, 10% (w/v) substrate concentration and 15 FPU/g substrate of commercial cellulase. The results indicate that it is possible to reach SSF yields in the range of 50–72% of the maximum theoretical SSF yield, based on the glucose available in pretreated materials, in 72–82 h. Maximum ethanol contents from 16 to 19 g/l were obtained in fermentation media, depending on the material tested.  相似文献   

19.
Water-hyacinth (Eichhornia crassipes) hemicellulose acid hydrolysate has been utilized as a substrate for ethanol production using Pichia stipitis NRRL Y-7124. Hydrolysate fermentability was considerable improved by boiling, and overliming up to pH 10.0 with solid Ca(OH)(2) in combination with sodium sulfite. The percent total sugar utilized and ethanol yield (Y(p/s)) for the untreated hydrolysate were 20.15+/-0.17% and 0.19+/-0.003 g(p) g(s)(-1), respectively, compared with 76.0+/-0.32% and 0.35 g(p) g(s)(-1), respectively for the treated material. The fermentation was very effective at an aeration rate of 0.02 v/v/m, temperature 30+/-0.2 degrees C and pH 6.0+/-0.2. However, the volumetric productivity (Q(p)) was still considerably less than observed in a simulated synthetic hydrolysate medium with a sugar composition similar to the hemicellulose acid hydrolysate. L-Arabinose was not fermented but assimilated. The presence of acetic acid in the hydrolysate decreased the ethanol yield and productivity considerably.  相似文献   

20.
Aims:  To compare the bacteriostatic and bactericidal activity of 13 chemotyped essential oils (EO) on 65 bacteria with varying sensitivity to antibiotics.
Methods and Results:  Fifty-five bacterial strains were tested with two methods used for evaluation of antimicrobial activity (CLSI recommendations): the agar dilution method and the time-killing curve method. EO containing aldehydes ( Cinnamomum verum bark and Cymbopogon citratus ), phenols ( Origanum compactum , Trachyspermum ammi , Thymus satureioides , Eugenia caryophyllus and Cinnamomum verum leaf) showed the highest antimicrobial activity with minimum inhibitory concentration (MIC) <2% (v/v) against all strains except Pseudomonas aeruginosa . Alcohol-based EO ( Melaleuca alternifolia , Cymbopogon martinii and Lavandula angustifolia ) exhibited varying degrees of activity depending on Gram status. EO containing 1·8-cineole and hydrocarbons ( Eucalyptus globulus , Melaleuca cajeputii and Citrus sinensis ) had MIC90% ≥ 10% (v/v). Against P. aeruginosa , only C. verum bark and O. compactum presented MIC ≤2% (v/v). Cinnamomum verum bark, O. compactum , T. satureioides , C. verum leaf and M. alternifolia were bactericidal against Staphylococcus aureus and Escherichia coli at concentrations ranging from to 0·31% to 10% (v/v) after 1 h of contact. Cinnamomum verum bark and O. compactum were bactericidal against P. aeruginosa within 5 min at concentrations <2% (v/v).
Conclusions:  Cinnamomum verum bark had the highest antimicrobial activity, particularly against resistant strains.
Significance and Impact of the Study:  Bacteriostatic and bactericidal activity of EO on nosocomial antibiotic-resistant strains.  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