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 640 毫秒
1.
A membrane bioreactor for production of nisin Z was constructed using Lactococcus lactis IO-1 in continuous culture using hydrolyzed sago starch as carbon source. A strategy used to enhance the productivity of nisin Z was to maintain the cells in a continuous growth at high cell concentration. This resulted in a volumetric productivity of nisin Z, as 50,000 IU l−1 h−1 using a cell concentration of 15 g l−1, 30°C, pH 5.5 and a dilution rate of 1.24 h−1. Adding 10 g l−1 YE and 2 g l−1 polypeptone, other inducers were unnecessary to maintain production of nisin. The operating conditions of the reactor removed nisin and lactate, thus minimizing their effects which allowed the maintenance of cells in continuous exponential growth phase mode with high metabolic activity.  相似文献   

2.
Factors affecting the production of the rare sugar l-xylulose from xylitol using resting cells were investigated. An E. coli BPT228 strain that recombinantly expresses a gene for xylitol dehydrogenase was used in the experiments. The ratio of xylitol to l-xylulose was three times lower in the cytoplasm than in the medium. The effects of pH, temperature, shaking speed, and initial xylitol concentration on l-xylulose production were investigated in shaking flasks using statistical experimental design methods. The highest production rates were found at high shaking speed and at high temperature (over 44°C). The optimal pH for both productivity and conversion was between 7.5 and 8.0, and the optimal xylitol concentration was in the range 250–350 g l−1. A specific productivity of 1.09 ± 0.10 g g−1 h−1 was achieved in a bioreactor. The response surface model based on the data from the shake flask experiments predicted the operation of the process in a bioreactor with reasonable accuracy.  相似文献   

3.
Maltose and yeast extract were the most favourable carbon and nitrogen sources for exopolysaccharide production by submerged culture of Shiraia bambusicola WZ-003, and initial maltose and yeast extract concentrations were at 30 and 3 g l−1, respectively. Plant oils could increase the mycelial growth and exopolysaccharide production in tested concentration. K+ and Mg2+ could enhance the mycelial growth and exopolysaccharide biosynthesis. The optimal cultivation temperature and initial pH were found to be 26°C and 6.0, respectively. Exopolysaccharide concentration reached 0.53 g l−1 in 15-l fermenter under optimal nutritional conditions.  相似文献   

4.
Plant gum as an elicitor for guggulsterone production in cell cultures of Commiphora wightii is reported for the first time. Guggulsterone production increased 2.4 fold in the cell cultures by gum Arabic (100 mg l−1), while mesquite gum elicited 2 fold. The cells treated with gum Arabic at 7th and 9th day accumulated enhanced guggulsterones within 24 h, which increased further up to 48 h and then declined. The cells treated at 9th day accumulated higher amount (218 μg l−1) of guggulsterones after 48 h of elicitation as compared to cells treated at 7th day (164 μg l−1). The optimized elicitation conditions were used in vessels of varying capacity where maximum yield of 285 μg l−1 of guggulsterones was recorded in 3 l shake flasks. These experiments enabled highest guggulsterones yield in a short duration of 11 days in cell cultures of C. wightii.  相似文献   

5.
Cell cultures of Commiphora wightii (Arnott.) Bhandari were grown in shake flasks and a bioreactor and an increase in guggulsterone accumulation up to 18 μg l−1 was recorded in cells grown in the production medium containing a combination of sucrose:glucose (4% total), precursors (phenylalanine, pyruvic acid, xylose, and sodium acetate), morphactin, and 2iP. A yield of 10 g l−1 biomass and ∼200 μg l−1 guggulsterone was recorded in a 3-l flask and in a 2-l stirred tank bioreactor compared with 6.6 g biomass and 67 μg l−1 guggulsterone in 250-ml flasks. Increased vessel size was correlated with increased biomass and guggulsterone accumulation. 2iP alone was not effective for biomass and guggulsterone accumulation in cell cultures of C. wightii.  相似文献   

6.
Production of lactic acid from glucose by immobilized cells of Lactococcus lactis IO-1 was investigated using cells that had been immobilized by either entrapment in beads of alginate or encapsulation in microcapsules of alginate membrane. The fermentation process was optimized in shake flasks using the Taguchi method and then further assessed in a production bioreactor. The bioreactor consisted of a packed bed of immobilized cells and its operation involved recycling of the broth through the bed. Both batch and continuous modes of operation of the reactor were investigated. Microencapsulation proved to be the better method of immobilization. For microencapsulated cells at immobilized cell concentration of 5.3 g l−1, the optimal production medium had the following initial concentrations of nutrients (g l−1): glucose 45, yeast extract 10, beef extract 10, peptone 7.5 and calcium chloride 10 at an initial pH of 6.85. Under these conditions, at 37 °C, the volumetric productivity of lactic acid in shake flasks was 1.8 g l−1 h−1. Use of a packed bed of encapsulated cells with recycle of the broth through the bed, increased the volumetric productivity to 4.5 g l−1 h−1. The packed bed could be used in repeated batch runs to produce lactic acid.  相似文献   

7.
A recombinant β-galactosidase from Sulfolobus solfataricus produced galactooligosaccharides (GOS) from lactose by transgalactosylation. The enzyme activity for GOS production was maximal at pH 6.0 and 85°C. The half-lives of the recombinant β-galactosidase at 70, 75, 80, 85, and 90°C were 700, 111, 72, 43, and 2.4 h, respectively, and its deactivation energy was 213 kJ mol−1. The optimal amount of enzyme for effective GOS production was 3.6 U of enzyme ml−1. GOS production increased with increasing lactose concentration, whereas the yield of GOS from lactose was almost constant. The rates of hydrolysis and transgalactosylation reactions increased with increasing temperature but the final concentration of GOS was maximal at 80°C. Under the conditions of pH 6.0, 80°C, 600 g lactose l−1, and 3.6 U enzyme ml−1, 315 g GOS l−1 were obtained for 56 h with a yield of 52.5% (w/w). The β-galactosidase from S. solfataricus produced GOS with the highest concentration and yield among thermostable β-galactosidases reported to date.  相似文献   

8.

Phenylethanoid glycosides (PeG) are a class of polyphenols found in some plants that have pharmaceutical effects as anti-inflammatories and anti-oxidants. The presence of PeG (acteoside) in the aerial parts of Scrophularia striata Boiss. has been demonstrated. Considerable progress has been made using plant cell cultures to stimulate formation and accumulation of secondary metabolites. The present study optimized phenylethanoid production from shake flasks to bioreactor using a cell culture of S. striata. The optimal conditions for production of cell biomass by scale-up to a bioreactor were determined to be a pH of 4.8, air flow rate of 0.5–1.5 l min−1, and mixing speed of 110–170 rpm at 25 ± 1 °C in darkness. Growth parameters and PeG production were measured and compared with the results from the shake flasks. The results showed that cell biomass was high in the bioreactor (15.64 g l−1 DW) and in the shake flasks (14.16 g l−1 DW). The acteoside content in the bioreactor was 1404.20 μg g−1 DW, which is threefold higher than in the shake flasks (459.71 μg g−1 DW). The echinacoside concentration in the bioreactor was 1449.39 μg g−1, 1.36-fold lower than in the shake flasks (1973.03 μg g−1 DW). This study established an efficient way for production of acteoside, the major PeG, in a bioreactor.

  相似文献   

9.
Paclitaxel and 10-deacetylbaccatin III (10-DAB III) were produced in suspension cultures of Taxus × media var. Hicksii grown in shake-flasks and in a 7-l bioreactor reaching, in the bioreactor, 4.4 mg l−1 (on day 14) and 37.5 mg l−1 (on day 11). In shake-flasks the highest total content of paclitaxel and 10-DAB III was 7.3 mg l−1 (on day 4) and 8.8 mg l−1 (on day 18). Phenylalanine, at 0.05 mM, increased paclitaxel accumulation in cells cultivated in bioreactor and flasks 30-fold and 9-fold (from 0.02 mg l−1 to 0.6 mg l−1 and to 0.2 mg l−1, respectively). The 10-DAB III content in cells from flasks was increased from 0.4 mg l−1 to 1.6 mg l−1.  相似文献   

10.
A mutant designated as UV-3 was obtained from wild-type Enterobacter aerogenes 10293 through u.v. radiation. The activities of α-acetolactate decarboxylase (Ald), lactate dehydrogenase (Ldh) and diacetyl reductase (Dr) in UV-3 were strongly attenuated, with the lowest activities at pH 7.0–7.5, and temperature between 36 and 39°C. Compared to the wild-type, the yield of diacetyl by UV-3 was increased 18.7-fold, up to 1.05 ± 0.01 g l−1. Acetoin and ethanol productions were decreased by 48.4 and 71.4%, respectively, but acetate yield was increased by 34.6%. Optimum medium for diacetyl production by UV-3 contained 10% glucose, 0.5% peptone, 0.5% yeast extract powder, 0.01% (NH4)2SO4, 0.1% citric acid, 0.2% MnSO4 and 0.2% MgSO4, and this was determined by one-factor-at-a-time approach. Data from the five level central composite designs demonstrated that initial pH of 7.0, temperature of 37°C and rotational speed of 180 rev/min were optimum processing parameters for diacetyl production. The maximum yield of diacetyl could reach 1.35 g l−1 in a 5-l bioreactor. These results showed an enhancement of the non-enzymatic oxidative decarboxylation of α-acetolactate and a decrease in the activities of Ald, Ldh and Dr as a consequence of diacetyl accumulation in UV-3.  相似文献   

11.
In this work, the production of 1,3-propanediol from glucose and molasses was studied in a two-step process using two recombinant microorganisms. The first step of the process is the conversion of glucose or other sugar into glycerol by the metabolic engineered Saccharomyces cerevisiae strain HC42 adapted to high (>200 g l−1) glucose concentrations. The second step, carried out in the same bioreactor, was performed by the engineered strain Clostridium acetobutylicum DG1 (pSPD5) that converts glycerol to 1,3-propanediol. This two-step strategy led to a flexible process, resulting in a 1,3-propanediol production and yield that depended on the initial sugar concentration. Below 56.2 g l−1 of sugar concentration, cultivation on molasses or glucose showed no significant differences. However, at higher molasses concentrations, glycerol initially produced by yeast could not be totally converted into 1,3-propanediol by C. acetobutylicum and a lower 1,3-propanediol overall yield was observed. In our hand, the best results were obtained with an initial glucose concentration of 103 g l−1, leading to a final 1,3-propanediol concentration of 25.5 g l−1, a productivity of 0.16 g l−1 h−1 and 1,3-propanediol yields of 0.56 g g−1 glycerol and 0.24 g g−1 sugar, which is the highest value reported for a two-step process. For an initial sugar concentration (from molasses) of 56.2 g l−1, 27.4 g l−1 of glycerol were produced, leading to 14.6 g l−1 of 1.3-propanediol and similar values of productivity, 0.15 g l−1 h−1, and overall yield, 0.26 g g−1 sugar.  相似文献   

12.
A new bacterial strain producing succinic acid was enriched from bovine rumen content. It is facultatively anaerobic, belongs to the family Pasteurellaceae and has similarity to the genus Mannheimia. In batch cultivations with D-glucose or sucrose the strain produced up to 5.8 g succinic acid l−1 with a productivity and a yield of up to 1.5 g l−1 h−1 and 0.6 g g−1, respectively. With crude glycerol up to 8.4 g l−1, 0.9 g l−1 h−1 and 1.2 g g−1 were obtained. Electronic supplementary material  The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

13.
Biomass and lipid productivities of Chlorella vulgaris under different growth conditions were investigated. While autotrophic growth did provide higher cellular lipid content (38%), the lipid productivity was much lower compared with those from heterotrophic growth with acetate, glucose, or glycerol. Optimal cell growth (2 g l−1) and lipid productivity (54 mg l−1 day−1) were attained using glucose at 1% (w/v) whereas higher concentrations were inhibitory. Growth of C. vulgaris on glycerol had a similar dose effects as those from glucose. Overall, C. vulgaris is mixotrophic.  相似文献   

14.
Although several reports have been published on recombinant protein expression using Drosophila cells, information on their metabolism and growth in vitro is relatively scarce. In the present study, we have analyzed the growth and metabolism of transfected S2 cells (S2AcRVGP) in bioreactor cultures with serum-free medium Sf900 II, to evaluate its potential for mass production of a rabies virus glycoprotein (RVGP). Cells were cultured in a 3 l-stirred-tank bioreactor at 28 °C with pH controlled at 6.2 and dissolved oxygen at 50% air saturation. The cells attained a specific growth rate and maximum cell density as high as 0.084 h−1 and 2.3 × 10cell ml−1, respectively. The main substrates consumed during this rapid growth phase were glucose, glutamine and proline. An atypical accumulation of ammonia and alanine was observed in the culture medium, up to 62 mM and 47 mM, respectively, but lactate was produced in low levels. After exhaustion of glutamine and proline as energy sources, alanine was consumed and production of ammonia increased. The production of recombinant RVGP reached concentrations as high as 178 μg l−1. Premature exhaustion of glutamine, serine and cysteine could be related to degradation of the recombinant glycoprotein. In general, the results demonstrated that S2AcRVGP can be considered an effective vehicle for large-scale recombinant glycoprotein expression and that several critical factors of the bioprocess could be optimized to increase the quality and productivity of the RVGP.  相似文献   

15.
The production of l-phenylalanine is conventionally carried out by fermentations that use glucose or sucrose as the carbon source. This work reports on the use of glycerol as an inexpensive and abundant sole carbon source for producing l-phenylalanine using the genetically modified bacterium Escherichia coli BL21(DE3). Fermentations were carried out at 37°C, pH 7.4, using a defined medium in a stirred tank bioreactor at various intensities of impeller agitation speeds (300–500 rpm corresponding to 0.97–1.62 m s−1 impeller tip speed) and aeration rates (2–8 L min−1, or 1–4 vvm). This highly aerobic fermentation required a good supply of oxygen, but intense agitation (impeller tip speed ~1.62 m s−1) reduced the biomass and l-phenylalanine productivity, possibly because of shear sensitivity of the recombinant bacterium. Production of l-phenylalanine was apparently strongly associated with growth. Under the best operating conditions (1.30 m s−1 impeller tip speed, 4 vvm aeration rate), the yield of l-phenylalanine on glycerol was 0.58 g g−1, or more than twice the best yield attainable on sucrose (0.25 g g−1). In the best case, the peak concentration of l-phenylalanine was 5.6 g L−1, or comparable to values attained in batch fermentations that use glucose or sucrose. The use of glycerol for the commercial production of l-phenylalanine with E. coli BL21(DE3) has the potential to substantially reduce the cost of production compared to sucrose- and glucose-based fermentations.  相似文献   

16.
The recombinant Pichia pastoris harboring an improved methionine adenosyltransferase (MAT) shuffled gene was employed to biosynthesize S-adenosyl-l-methionine (SAM). Two l-methionine (l-Met) addition strategies were used to supply the precursor: the batch addition strategy (l-Met was added separately at three time points) and the continuous feeding strategies (l-Met was fed continuously at the rate of 0.1, 0.2, and 0.5 g l−1 h−1, respectively). SAM accumulation, l-Met conversion rate, and SAM productivity with the continuous feeding strategies were all improved over the batch addition strategy, which reached 8.46 ± 0.31 g l−1, 41.7 ± 1.4%, and 0.18 ± 0.01 g l−1 h−1 with the best continuous feeding strategy (0.2 g l−1 h−1), respectively. The bottleneck for SAM production with the low l-Met feeding rate (0.1 g L−1 h−1) was the insufficient l-Met supply. The analysis of the key enzyme activities indicated that the tricarboxylic acid cycle and glycolytic pathway were reduced with the increasing l-Met feeding rate, which decreased the adenosine triphosphate (ATP) synthesis. The MAT activity also decreased as the l-Met feeding rate rose. The reduced ATP synthesis and MAT activity were probably the reason for the low SAM accumulation when the l-Met feeding rate reached 0.5 g l−1 h−1.  相似文献   

17.
The green microalga Chlorella zofingiensis can produce the ketocarotenoid astaxanthin under heterotrophic culture conditions. Here we report the growth-associated biosynthesis of astaxanthin in this biotechnologically important alga. With glucose as sole carbon and energy source, C. zofinginesis grew fast in the dark with rapid exhaustion of nitrogen and carbon sources from media, leading to a high specific growth rate (0.034 h−1). Cultures started at a cell concentration of about 3.4 × 109 cells l−1 reached, after 6 days, standing biomass values of 1.6 × 1011 cells or 8.5 g dry weight l−1. Surprisingly, the biosynthesis of astaxanthin was found to start at early exponential phase, independent of cessation of cell division. A general trend was observed that the culture conditions benefiting cell growth also benefited astaxanthin accumulation, indicating that astaxanthin was a growth-associated product in this alga. The maximum cell dry biomass and astaxanthin yield were 11.75 g l−1 and 11.14 mg l−1 (about 1 mg g−1), simultaneously obtained in the fed-batch culture with a combined glucose–nitrate mixture addition, which were the highest ever reported in dark-heterotrophic algal cultures. The possible reasons why dark-heterotrophic C. zofingiensis could produce astaxanthin during the course of cell growth were discussed.  相似文献   

18.
Cui  Fengjie  Li  Yin  Liu  Zhiqiang  Zhao  Hui  Ping  Lifeng  Ping  Liying  Yang  Yinan  Xue  Yaping  Yan  Lijiao 《World journal of microbiology & biotechnology》2009,25(4):721-725
The objective of this study was to maximize production of xylanase by a newly isolated strain Penicillium thiersii ZH-19. Response surface methodology was employed to study the effects of significant factors such as pH, temperature, xylan concentration, and cultivation time, on the production of xylanase by Penicillium thiersii ZH-19. The optimal fermentation parameters for enhanced xylanase production were found to be pH 7.72, temperature 24.8°C, xylan 13.2 g l−1 and the fermentation time 125.8 h. The model predicted a xylanase activity of 75.24 U ml−1. Verification of the optimization showed that the maximum xylanase production reached 73.50 U mL−1 in the flask experiments and 80.23 U mL−1 in the scale of 15-L fermenter under the optimal condition.  相似文献   

19.
The present study describes a protocol for plant regeneration via somatic embryogenesis in temporary immersion system (TIS) for Camptotheca acuminata. Somatic embryos were induced by culturing hypocotyl segments from 14-day-old in vitro grown C. acuminata seedlings in TIS. Hypocotyl segments were placed in culture vessels modified with a mechanical device to support the fixation of explants. Cultures were maintained under a 16 h photoperiod with a light intensity of 60 μmol m−2 s−1 PPF at 25 ± 1°C. After 16 weeks of incubation embryogenic calli were formed above the edge of the mechanical device in the basal Murashige and Skoog (MS) medium containing 35 g l−1 sucrose and without hormonal supplementation. For plantlet regeneration, somatic embryos at cotyledonary stage were cultured in three different concentrations of 6-benzylamino-purine (0.5, 1.0 and 1.5 mg l−1 BAP) and in plant growth regulator (PGR) free medium. In general, 0.5 mg l−1 BAP was found to be the most effective concentration for growth and development of Camptotheca embryos in TIS. Conversion of somatic embryos into plantlets was also successfully achieved on sterile substrates moistened with 0.5 mg l−1 BAP. Plantlets derived from cotyledonary embryos were rooted in vitro with 0.5 mg l−1 indole-3-butyric acid (IBA) before transfer to ex vitro conditions.  相似文献   

20.
Covalent immobilization of Aspergillus awamori NRRL 3112 was conducted onto modified polyacrylonitrile membrane with glutaraldehyde as a coupling agent. The polymer carrier was preliminarily modified in an aqueous solution of NaOH and 1,2-diaminoethane. The content of amino groups was determined to be 0.58 mgeq g−1. Two ways of immobilization were used—in the presence of 0.2 g l−1 phenol and without phenol. The capability of two immobilized system to degrade phenol (concentration—0.5 g l−1) as a sole carbon and energy source was investigated in batch experiments. Seven cycles of phenol biodegradation were conducted. Better results were obtained with the immobilized system prepared in the presence of phenol, regarding degradation time and phenol biodegradation rate. Scanning electron micrographs of the polyacrylonitrile membrane/immobilized Aspergillus awamori NRRL at the beginning of repeated batch cultivation and after the 7th cycle were compared. After the 7th cycle of cultivation the observations showed large groups of cells. The results from the batch experiments with immobilized system were compared to the results produced by the free strain. Phenol biodegradation experiments were carried out also in a bioreactor with spirally wound membrane with bound Aspergillus awamori NRRL 3112 in a regime of recirculation. 10 cycles of 0.5 g l−1 phenol biodegradation were run consecutively to determine the degradation time and rate for each cycle. The design of the bioreactor appeared to be quite effective, providing large membrane surface to bind the strain.  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