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.
Bacillus subtilis CBTK 106, isolated from banana wastes, produced high titres of α-amylase when banana fruit stalk was used as substrate in a solid-state fermentation system. The effects of initial moisture content, particle size, cooking time and temperature, pH, incubation temperature, additional nutrients, inoculum size and incubation period on the production of α-amylase were characterised. A maximum yield of 5 345 000 U mg-1 min-1 was recorded when pretreated banana fruit stalk (autoclaved at 121 °C for 60 min) was used as substrate with 70% initial moisture content, 400 μm particle size, an initial pH of 7.0, a temperature of 35 °C, and additional nutrients (ammonium sulphate/sodium nitrate at 1.0%, beef extract/peptone at 0.5%, glucose/sucrose/starch/maltose at 0.1% and potassium chloride/sodium chloride at 1.0%) in the medium, with an inoculum-to-substrate ratio of 10% (v/w) for 24 h. The enzyme yield was 2.65-fold higher with banana fruit stalk medium compared to wheat bran. Received: 18 April 1995/Received last revison: 6 May 1996/Accepted: 9 May 1996  相似文献   

2.
 According to their ability to synthesize 1,3-propanediol from glycerol, two species were isolated from the anoxic mud of a distillery waste-water digestor: Clostridium butyricum and Enterobacter agglomerans. The latter, a facultatively anaerobic gram-negative bacterium, is described for the first time as a microorganism producing 1,3-propanediol from glycerol. The products of glycerol conversion by E. agglomerans were identified using nuclear magnetic resonance. A 20-g/l glycerol solution was fermented mainly to 1,3-propanediol (0.51 mol/mol) and acetate (0.18 mol/mol). Ethanol, formate, lactate and succinate were formed as by-products. Gas production was very low; 1,3-propanediol production perfectly balanced the oxido-reduction state of the microorganism. Acetate was the predominant metabolite generating energy for growth. High-glycerol-concentration fermentations (71 g/l and 100 g/l) resulted in an increase of the 1,3-propanediol yield (0.61 mol/mol) at the expense of lactate and ethanol production. Specific rates of glycerol consumption and 1, 3-propanediol and acetate production increased whereas the growth rate decreased. The decrease in ATP yield was linearly correlated with the specific rate of 1,3-propanediol production. Incomplete glycerol consumption (about 40 g/l) was systematically observed when high glycerol concentrations were used. The unbalanced oxido-reduction state, the low carbon recovery and the detection of an unknown compound by HPLC observed in these cases indicate the formation of another metabolite, which is possibly an inhibitory factor. Received: 17 November 1994 / Accepted: 15 December 1994  相似文献   

3.
 The objective of this study was to assess fermentation product, growth rate and growth yield responses of Selenomonas ruminantium HD4 to limiting and non-limiting ammonia concentrations. The ammonia half-inhibition constant for S. ruminantium in batch culture was 296 mM. Cells were grown in continuous culture with a defined ascorbate-reduced basal medium containing either 0.5, 5, 25, 50, 100 or 200 mM NH4Cl and dilution rates were 0.07, 0.14, 0.24 or 0.40 h-1. Ammonia was the growth-limiting nutrient when 0.5 mM NH4Cl was provided and the half-saturation constant was 72 μM. Specific rates of glucose utilization and fermentation acid carbon formation were highest for 0.5 mM NH4Cl. Lactate production (moles per mole of glucose disappearing) increased at the fastest dilution rate (0.40 h-1) for 5.0 mM NH4Cl while acetate and propionate decreased when compared to slower dilutions (0.07 and 0.14 h-1). Lactate production remained low while acetate and propionate remained high for all dilution rates when NH4Cl concentrations were 25 mM or greater. Yield (Y Glc and Y ATP) were nearly doubled when NH4Cl was increased from 0.5 mM (25.1 g cells/mol glucose used and 13.9 g cells/mol ATP produced respectively) to the higher concentrations. Y Glc was highest at 25 mM and 50 mM NH4Cl (48.2 cells/mol and 43.1 cells/mol respectively) as was Y ATP (23.2 cells/mol and 20.8 cells/mol respectively). Y NH3 was highest at the lowest NH4Cl concentration. The maximal fermentation product formation rate occurred at a growth-limiting ammonia concentration, while maximal glucose and ATP bacterial yields occurred at non-growth-limiting ammonia concentrations. Given the growth response of this ruminal bacterium, it is possible that maximization of ruminal bacterial yield may necessitate sacrificing the substrate degradation rate and vice versa. Received: 5 December 1995/Received revision: 2 April 1996/Accepted: 22 April 1996  相似文献   

4.
Candida shehatae were sequentially subjected to aerobic conditions for cellular growth, followed by anaerobic conditions for ethanol production from D-xylose at pH 2.5, 4.5 and 6.0. Ethanol yields increased from 0.25 g/g to 0.37 g/g and xylitol yields decreased from 0.33 g/g to 0.1 g/g as the pH was increased from 2.5 to 6.0. Cell viability, measured by plate counts and methylene blue staining, decreased in all of the fermentations, following the switch from aerobic to anaerobic conditions. However, pH 6.0 was shown to extend cell viability and increase the final ethanol concentration from 45 g/l to 55 g/l, compared to the yield at pH 4.5. Received: 25 April 1995/Received revision: 5 September 1995/Accepted: 20 September 1995  相似文献   

5.
Instead of the conventional carbon sources used for propionic acid biosynthesis, the utilization of glycerol is considered here, since the metabolic pathway involved in the conversion of glycerol to propionic acid is redox-neutral and energetic. Three strains, Propionibacterium acidipropionici, Propionibacterium acnes and Clostridium propionicum were tested for their ability to convert glycerol to propionic acid during batch fermentation with initially 20 g/l glycerol. P. acidipropionici showed higher efficiency in terms of fermentation time and conversion yield than did the other strains. The fermentation profile of this bacterium consisted in propionic acid as the major product (0.844 mol/mol), and in minimal by-products: succinic (0.055 mol/mol), acetic (0.023 mol/mol) and formic (0.020 mol/mol) acids and n-propanol (0.036 mol/mol). The overall propionic acid productivity was 0.18 g l−1h−1. A comparative study with glucose and lactic acid as carbon sources showed both less diversity in end-product composition and a 17% and 13% lower propionic acid conversion yield respectively than with glycerol. Increasing the initial glycerol concentration resulted in an enhanced productivity up to 0.36 g l−1h−1 and in a maximal propionic acid concentration of 42 g/l, while a slight decrease of the conversion yield was noticed. Such a propionic acid production rate was similar or higher than the values obtained with lactic acid (0.35 g l−1h−1) or glucose (0.28 g l−1h−1). These results demonstrated that glycerol is a carbon source of interest for propionic acid production. Received: 15 July 1996 / Received revision: 11 November 1996 / Accepted: 11 November 1996  相似文献   

6.
 The synthesis of poly(3-hydroxyalkanoates) (PHA) by Pseudomonas putida KT2442 growing on long-chain fatty acids was studied in continuous cultures. The effects of the growth rate on the biomass and polymer concentration were determined and it was found that the PHA concentrations decreased with increasing growth rates. The highest volumetric productivity was 0.13 g PHA l-1 h-1 at a specific growth rate (μ) of 0.1 h-1. The molecular mass of the polymer remained constant at all growth rates but changes in the monomeric composition of the PHA synthesized were observed. Variation of the carbon to nitrogen (C/N) ratio of the substrate feed at μ=0.1 h-1 revealed optimal PHA formation at C/N=20 mol/mol. In order to optimize PHA production P. putida KT2442 was cultivated to high cell densities in oxygen-limited continuous cultures. In this way a maximum biomass concentration of 30 g/l containing approximately 23% PHA was achieved. This corresponds to a volumetric productivity of 0.69 g  l-1 h-1. Received: 14 December 1995 / Received revision: 18 April 1996 / Accepted: 22 April 1996  相似文献   

7.
Degradation of pyrene by Mycobacterium flavescens   总被引:1,自引:0,他引:1  
 A strain of Mycobacterium flavescens was isolated from polluted sediments. It was capable of utilizing pyrene as a sole source of carbon and energy. When pyrene was supplied as a suspension at 50 μg/ml, the generation time was 9.6 h and the rate of pyrene utilization was 0.56 μg ml-1 day-1. In addition to pyrene, the strain could mineralize phenanthrene (17.7%) and fluoranthene (17.9%), but failed to mineralize naphthalene, chrysene, anthracene, fluorene, acenaphthene and benzo[a]pyrene, as determined by recovery of radiolabeled CO2 in incubations conducted for 2 weeks under growth conditions. Metabolites produced during growth on pyrene were detected and characterized by HPLC and GC-MS. The product of initial ring oxidation, 4,5-dihydroxy-4,5-dihydropyrene was identified, as well as ring-fission products including 4-phenanthroic acid, phthalic acid, and 4,5-phenanthrenedioic acid. Received: 3 October 1995/Received last revision: 1 April 1996/Accepted: 15 April 1996  相似文献   

8.
 The thermotolerant, ethanol-producing yeast strain, Kluyveromyces marxianus IMB3, was shown to produce ethanol at 45°C on starch-containing media supplemented with a crude amylase preparation derived from the thermophilic, filamentous fungus Talaromyces emersonii CBS 813.70. Ethanol production on media containing 4% (w/v) starch increased to a maximum of 15 g/l with 40 h, and this represented 74% of the maximum theoretical yield. Subsequent experimentation involving growth of both organisms in fermentations on starch-containing media (4% w/v) demonstrated that the mixed-culture system was capable of ethanol production at 45°C with maximum yields at 12 g/l obtained with 65 h. The advantages associated with ethanol production by this system are discussed. Received: 16 May 1994/Accepted: 22 October 1994  相似文献   

9.
 Previously it was demonstrated that bacteria are capable of transforming soluble uranyl ion, U(VI), to insoluble uraninite, U(IV); however, the rate for this transformation has not been determined. We report the kinetic coefficients for Desulfovibrio desulfuricans DSM 1924 grown in a continuous-flow chemostat where pyruvate was the electron donor and sulfate was the electron acceptor. The medium was supplemented with 1 mM uranyl nitrate, and the chemostat flow rate ranged from 1.12 ml/h to 4.75 ml/h with incubation at 28°C. The maximum rate of pyruvate utilization (k) was determined to be 4.7 days-1, while the half-velocity constant (K s) was 127 mg/l. The yield coefficient (Y) of cells per mole of pyruvate oxidized was calculated to be 0.021 g, while the endogenous decay coefficient (k d) was determined to be 0.072 days-1. More than 90% of U(VI) was transformed to U(VI) in the chemostat under the conditions employed. Received: 7 September 1995/Received last revision: 10 January 1996/Accepted: 5 February 1996  相似文献   

10.
 A continuous bioreactor packed with a fibrous matrix was set up. Cells of Pediococcus acidilactici PO2 were inoculated and MRS broth was fed gradually until cell growth and immobilization were achieved. Kinetics of fermentation and production of bacteriocin were investigated at dilution rates ranging from 0.63 day-1 to 1.58 day-1 and at pH values that varied between 4.0 and 5.5. A maximum bacteriocin activity of 6400 AU/ml was detected when the medium was fermented at dilution rates of at least 1.19 day-1 and the pH controlled at 4.5. The maximum bacteriocin productivity was 1.0×107 AUl-1 day-1 at a dilution rate of 1.58 day-1 and pH 4.5. At this high dilution rate, 1.21 g cells/l medium was produced, 95.9% of the glucose in MRS broth was utilized, and 15.1 g lactic acid/l accumulated in the bioreactor effluent. The bioreactor was operated continuously for 3 months without encountering any clogging, degeneration, or contamination problems, indicating good long-term stability of the bioreactor for bacteriocin production. About 94% of the cells in the bioreactor were immobilized, and the remainder were suspended in the medium. According to scanning electron microscopic observations, cell immobilization in the fibrous matrix was attained by natural attachment to fiber surfaces and entrapment in the void volume within the fibrous matrix. In conclusion, conditions for the optimum continuous production of pediocin were defined; this may facilitate the development of large-scale industrial processes for production of this bacteriocin. Received: 25 September 1995/Received revision: 30 November 1995/Accepted: January 1996  相似文献   

11.
 The Zygomycete Cunninghamella elegans produces the polycarboxylate siderophore rhizoferrin. Production depends mainly on iron concentration in the medium. With an optimized production medium the yield of rhizoferrin in a bioreactor could be increased to more than 4 g/l. Supplementation of the basic production medium with different precursors led to the formation of nine new rhizoferrins. Both the diaminobutane backbone and the citric acid side-chains of rhizoferrin could be substituted by appropriate analogues. These substitutions led to new siderophores either with a variable length of diamine bridge or with fewer or different functional groups. The proportion of the new diamine analogues relative to the total rhizoferrin could be markedly increased by the use of α-difluoromethylornithine, an inhibitor of the ornithine decarboxylase. Received: 11 September 1995/Received revision: 15 January 1996/Accepted: 22 January 1996  相似文献   

12.
 In order to determine the possible effect of nutrient limitations on the response of Corynebacterium glutamicum to a saline osmotic up-shock, the bacteria were grown in continuous cultures, at osmotic pressures of 0.4 osmol/kg and 1.2 osmol/kg, under ammonia and potassium limitation. At the low osmolality of 0.4 osmol/kg, the glutamate and proline levels of 15 mg/g and 5 mg/g dry weight respectively were lower than previously reported in glucose-limited continuous cultures (50 mg/g and 10 mg/g dry weight respectively). On the other hand, the internal trehalose pool was much higher at 40 mg/g dry weight. When the medium osmolality was increased to 1.2 osmol/kg by NaCl addition, under ammonia limitation, the proline content rose from 5 mg/g to 20 mg/g dry weight and the trehalose content from 40 mg/g to 70 mg/g dry weight, whereas the intracellular pool of glutamate remained essentially constant. An increase in the internal sodium content was also observed. Similar results were found for the internal pool of glutamate, proline and trehalose when C. glutamicum was grown under potassium limitations at an osmolality of 1.2 osmol/kg. There were also higher levels of sodium ions, glutamine and alanine. According to the present results, whereas proline was previously reported to be the dominantly accumulated osmoprotectant in C. glutamicum grown under glucose limitations, under ammonia and potassium limitations trehalose represented the dominantly synthesized metabolite. Received: 19 December 1995/Received revision: 9 April 1996/Accepted: 15 April 1996  相似文献   

13.
 The strain Penicillium purpurogenum P-26 was subjected to UV irradiation and N-methyl-N′-nitro-N-nitrosoguanidine treatment and mutants were isolated capable of synthesizing cellulase under the conditions of a high concentration of glucose. Initially mutants resistant to catabolite repression by 2-deoxy-D-glucose were isolated on Walseth’s cellulose/agar plates containing 15–45 mM 2-deoxy-D-glucose. These mutants were again screened for resistance to catabolite repression by glycerol or glucose on Walseth’s cellulose/agar plates containing 50 g/l glycerol or 50 g/l glucose respectively. Four mutants with different sizes of clearing zone on Walseth’s cellulose/agar plates containing 50 g/l glucose were selected for flask culture. Among them, the mutant NTUV-45-4 showed better carboxymethylcellulase activity in flask culture containing 1% Avicel plus 3% glucose than did the parental strain. Received: 9 October 1995/Received revision: 27 November 1995/Accepted: 8 January 1996  相似文献   

14.
Disposable sensor for biochemical oxygen demand   总被引:6,自引:0,他引:6  
 Disposable-type microbial sensors were prepared for the determination of biochemical oxygen demand (BOD). The yeast, Trichosporon cutaneum, was directly immobilized on the surface of miniature oxygen electrodes using an ultraviolet crosslinking resin (ENT-3400). The oxygen electrodes (15 mm× 2 mm×0.4 mm) were made on silicon substrates using micromachining techniques. They were Clark-type two-electrode systems with−1021 mV applied to the working electrode. Typical response times of the BOD sensors were in the range of 7–20 min. At 20°C, the sensors’ dynamic range was from 0 to 18 mg/l BOD when a glucose/glutamate BOD standard solution was used. The lower limit of detection was 0.2 mg/l BOD. This value was one order of magnitude lower than that of sensors previously reported. The sensors’ operational lifetime of 3 days was satisfactory for a disposable type. The sensors’ responses were reproducible to within 8% relative standard deviation. The BOD sensors’ were applied to untreated and treated waste waters from industrial effluents and municipal sewage. BOD values determined using these sensors correlated well with those determined by the conventional 5-day BOD determination method. Received: 22 December 1995/Received revision: 19 February 1996/Accepted: 17 March 1996  相似文献   

15.
Glycerol formation is vital for reoxidation of nicotinamide adenine dinucleotide (reduced form; NADH) under anaerobic conditions and for the hyperosmotic stress response in the yeast Saccharomyces cerevisiae. However, relatively few studies have been made on hyperosmotic stress under anaerobic conditions. To study the combined effect of salt stress and anaerobic conditions, industrial and laboratory strains of S. cerevisiae were grown anaerobically on glucose in batch-cultures containing 40 g/l NaCl. The time needed for complete glucose conversion increased considerably, and the specific growth rates decreased by 80–90% when the cells were subjected to the hyperosmotic conditions. This was accompanied by an increased yield of glycerol and other by-products and reduced biomass yield in all strains. The slowest fermenting strain doubled its glycerol yield (from 0.072 to 0.148 g/g glucose) and a nearly fivefold increase in acetate formation was seen. In more tolerant strains, a lower increase was seen in the glycerol and in the acetate, succinate and pyruvate yields. Additionally, the NADH-producing pathway from acetaldehyde to acetate was analysed by overexpressing the stress-induced gene ALD3. However, this had no or very marginal effect on the acetate and glycerol yields. In the control experiments, the production of NADH from known sources well matched the glycerol formation. This was not the case for the salt stress experiments in which the production of NADH from known sources was insufficient to explain the formed glycerol.  相似文献   

16.
 Hollow-fibre modules containing microporous membrane material were evaluated as bioreactors for waste gas treatment. The reactors were inoculated with the propene-utilizing strain Xanthobacter Py2, which formed a biofilm on the inner side of the fibres. The removal of the poorly soluble volatile propene from synthetic waste gas was monitored for up to 170 days. The maximum removal rates were 70–110 g propene per m3 reactor per hour. A gas residence time of 80 s was required to remove 95% of an initial propene concentration of 0.84 g/m3. The presence of ammonium in the liquid medium resulted in the development of an additional population of nitrifying organisms. Therefore, nitrate was used as the source of nitrogen in later experiments. During long-term operation, the propene removal rates gradually decreased. At low liquid velocities (1–5 cm/s) clogging of individual fibres with excess biomass was observed. Elevation of the liquid velocity in the fibres to 90 cm/s resulted in the formation of a dense biofilm and prevented clogging of the fibres. However, also at this high liquid velocity a gradual decrease in propene removal rate was observed. These results suggest that aging of biofilms is a very important factor in long-term operation of hollow-fibre bioreactors. Received: 24 November 1995 / Received revision: 14 February 1996 / Accepted: 20 February 1996  相似文献   

17.
不同发酵条件下产甘油假丝酵母有机酸代谢的研究   总被引:3,自引:0,他引:3  
产甘油假丝酵母 (Candidaglycerolgenesis)发酵产生的有机酸对丙三醇产品质量和产率均有影响。发现在发酵其它条件恒定 ,装液比和玉米浆浓度增加时 ,发酵液总酸是递增的。在装液比为 0 2和玉米浆浓度为 8g L时 ,丙酮酸和乳酸在细胞生长期可分别积累达 4 1g L和 1 0g L ,比正常发酵时增加 2倍以上 ,丙三醇产率也低 ;然而 ,装液比为 0 0 8和玉米浆浓度为 4g L时 ,丙酮酸和乳酸产生较低 ,丙三醇产率较高 ,但乙酸积累比供氧不足时高 ,可达 2 6g L。发酵过程中有机酸被细胞代谢 ,含量逐渐下降 ,如在初糖浓度为 1 0 0g L时 ,有机酸在细胞生长期积累至高峰后 ,丙三醇和有机酸随之均降低至较低含量 ,并且丙酮酸或乳酸可以转化为乙酸。此外 ,在外加一些添加剂时对其产生有机酸也有影响 ,如添加 1 %油酸和VB1时可以降低乙酸的积累 ,同时增加丙酮酸的含量 ,丙三醇产量也有所增加 ;而丙酮酸结构类似物氟代丙酮酸和亚硫酸盐促进乙酸的产生 ,使酮戊二酸合成减少 ,丙三醇产量约增加 2 0 %。  相似文献   

18.
 Laccase of Trametes versicolor was generally able to oxidize anthracene in vitro. After 72 h incubation about 35% of the anthracene was transformed stoichiometrically to 9,10-anthraquinone. Transformation of anthracene increased rapidly in the presence of different mediators that readily generate stable radicals: 2,2′-azino-bis-(3-ethylbenzthiazoline-6-sulfonic acid) (ABTS) and 1-hydroxybenzotriazole. For the reaction, the presence of both the laccase and the mediator was necessary. In the presence of 0.005 mM 1-hydroxybenzotriazole this conversion had removed 47% of the anthracene after 72 h; 75% of the substrate was oxidized during this period when ABTS (1 mM) was used as mediator. In contrast to reactions without or with only low concentrations of a mediator, there was a discrepancy between the disappearance of anthracene and the formation of 9,10-anthraquinone in mediator-forced reactions. Coupling-products of mediators with anthracene degradation products were found. Anthracene disappeared nearly completely after incubation for 72 h with laccase in a 0.1 mM solution of 1-hydroxybenzotriazole and was transformed to 9,10-anthraquinone in about 80% yield; 90% of the substrate was transformed in the presence of ABTS (2.0 mM) resulting again in 80% quinone. Phenothiazine was not effective in this system. Received: 9 April 1996/Received last revision: 7 June 1996/Accepted: 10 June 1996  相似文献   

19.
 Eight strains of the genus Aureobasidium obtained from culture collections were tested for their capability to produce poly(β-L-malic acid) (PMA). Four of the tested strains showed positive results. The most productive strain, A. pullulans CBS 591.75, was used to study the production of PMA in stirred-tank reactors. It was found that PMA was mainly produced in the late exponential phase, and the production related positively to glucose consumption. At the beginning of the fermentation the pH increased from 4.0 to about 7.0; subsequently the pH decreased and remained stable at around 3.0–3.5 for several days. Temperatures higher than 25°C were detrimental to PMA production and cell growth. PMA production and cell growth at 20°C and 25°C exhibited no significant differences. PMA production and cell growth were studied under pH-controlled fermentation (at pH 2.0, 4.0, 5.5). The highest PMA production occurred at pH 4.0. PMA production was reduced at pH 2.0 although quite reasonable cell growth occurred at this pH value. Under optimized conditions 9.8 g PMA/l was produced during 9 days of fermentation in the stirred-tank reactors with an overall yield of 0.11 g PMA/g glucose. A procedure for the isolation of PMA and its separation from the other components of the fermentation broth was developed. The isolated PMA was characterized by 1H and 13C-NMR spectroscopy as well as by infrared absorption spectroscopy. Gel-permeation chromatography revealed a relative molecular mass of approximately 3000–5000 by comparison with polyethylene glycol standards. Received: 13 February 1996/Received revision: 25 April 1996/Accepted: 1 May 1996  相似文献   

20.
Cryptococcus curvatus is a yeast with industrial potential because it can grow and accumulate lipid on a very broad range of substrates. In this study we describe growth and lipid accumulation on glycerol in a fed-batch fermentation mode. We performed a fermentation consisting of two phases. The first phase is the biomass production phase in which there is no nutrient limitation except for very short periods of glycerol exhaustion. The substrate feed was controlled by the dissolved oxygen tension. In the second phase nitrogen limitation was introduced, which causes lipid accumulation. This way very high cell densities of 118 g/l in a 50-h fermentation could be reached. With a lipid production rate of 0.59 g lipid l-1h-1, a cellular lipid content of 25% was obtained. The growth and lipid accumulation phase are characterized by different cellular fatty acid compositions. In the growth phase, a relatively high amount of C18:2 (linoleic acid) is present, which is a major component of membrane lipids. C18:0 (stearic acid) and C18:1 (oleic acid) are major constituents of the accumulated triglycerides and therefore the relative amount of C18:2 decreases during the lipid accumulation phase. Received: 19 September 1995/Received revision: 28 December 1995/Accepted: 8 January 1996  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