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 62 毫秒
1.
Maximal activity of the immobilized d-psicose 3-epimerase from Agrobacterium tumefaciens on Duolite A568 beads was achieved at pH 9.0 and 55 °C with borate, and at pH 8.5 and 50 °C without borate. The half-lives of the immobilized enzyme at 50 °C with and without borate were increased 4.2- and 128-fold compared to that of the free enzyme without borate, respectively. The immobilized enzyme with borate produced 441 g l?1 psicose from 700 g l?1 fructose at pH 9.0 and 55 °C, whereas 193 g l?1 psicose was produced without borate at pH 8.5 and 50 °C after 120 min in a batch reaction. The immobilized enzyme in a packed-bed bioreactor without borate was produced continuously 325 g l?1 psicose from 500 g l?1 fructose at a dilution rate of 1.62 h?1 over a 236 h period with productivity of 527 g l?1 h?1 while that without borate produced 146 g l?1 psicose at 4.15 h?1 over a 384-h period with productivity of 606 g l?1 h?1. The operational half-lives of the enzyme with and without borate in the bioreactor were 601 and 645 h, respectively. In the present study, psicose was produced stably with high productivity using the immobilized d-psicose 3-epimerase in the presence of borate.  相似文献   

2.
Olive stones are an agro-industrial by-product abundant in the Mediterranean area that is regarded as a potential lignocellulosic feedstock for sugar production. Statistical modeling of dilute-sulphuric acid hydrolysis of olive stones has been performed using a response surface methodology, with treatment temperature and process time as factors, to optimize the hydrolysis conditions aiming to attain maximum d-xylose extraction from hemicelluloses. Thus, solid yield and composition of solid and liquid phases were assessed by empirical modeling. The highest yield of d-xylose was found at a temperature of 195 °C for 5 min. Under these conditions, 89.7% of the total d-xylose was recovered from raw material. The resulting solids from optimal conditions were assayed as substrate for enzymatic hydrolysis, while fermentability of hemicellulosic hydrolysates was tested using the d-xylose-fermenting yeast Pachysolen tannophilus. Both bioprocesses were considerably influenced by enzyme loading and inoculum size. In the enzymatic hydrolysis step, about 56% of cellulose was converted into d-glucose by using an enzyme/solid ratio of 40 FPU g−1, while in the fermentation carried out with a cell concentration of 2 g L−1 a yield of 0.44 g xylitol/g d-xylose and a global volumetric productivity of 0.11 g L−1 h−1 were achieved.  相似文献   

3.
In this work, an integrated one-step alkaline–extrusion process was tested as pretreatment for sugar production from barley straw (BS) biomass. The influence of extrusion temperature (T) and the ratio NaOH/BS dry matter (w/w) (R) into the extruder on pretreatment effectiveness was investigated in a twin-screw extruder at bench scale. A 23 factorial design of experiments was used to analyze the effect of process conditions [T: 50–100 °C; R: 2.5–7.5% (w/w)] on composition and enzymatic digestibility of pretreated substrate (extrudate). The optimum conditions for a maximum glucan to glucose conversion were determined to be R = 6% and T = 68 °C. At these conditions, glucan yield reached close to 90% of theoretical, while xylan conversion was 71% of theoretical. These values are 5 and 9 times higher than that of the untreated material, which supports the great potential of this one-step combined pre-treatment technology for sugar production from lignocellulosic substrates. The absence of sugar degradation products is a relevant advantage over other traditional methods for a biomass to ethanol production process since inhibitory effect of such product on sugar fermentation would be prevented.  相似文献   

4.
2,2-Bis(hydroxymethyl)butyric acid (BHMB) is an important multifunctional chemical for the emerging bio-based polymer industry. It can be produced from trimethylolpropane (TMP) by selective oxidation using growing cells of Corynebacterium sp. ATCC 21245. However, this process is limited by the low volumetric productivity and low concentration of the final product. In the present study, we performed sequential batch operation with cell recycling in media containing glycerol, acetic acid, and increasing concentrations of yeast extract. This approach enhanced the conversion of 10 and 15 g/L TMP to 11.0 and 16.3 g/L BHMB at rates of 0.50 and 0.20 g/L.h, respectively. Applying a cell bleeding strategy resulted in an overall 10-fold improvement in productivity. The consequently prolonged biocatalyst viability resulted in a quantitative conversion of 20 g/L TMP to 22.3 g/L BHMB and a yield of 1.10 gBHMB/gTMP (100% molar yield). This work facilitates further studies of the selective oxidation on other industrially important polyols.  相似文献   

5.
This paper reports development and implementation of superior fermentation strategies for β-galactosidase production by Lactobacillus acidophilus in a stirred-tank bioreactor. Process parameters (aeration and agitation) were optimized for the process by application of Central Composite Design. Aeration rate of 0.5 vvm and agitation speed of 250 rpm were most suitable for β-galactosidase production (2001.2 U/L). Further improvement of the operation in pH controlled environment resulted in 2135 U/L of β-galactosidase with productivity of 142.39 U/L h. Kinetic modeling for biomass and enzyme production and substrate utilization were carried out at the aforementioned pH controlled conditions. The logistic regression model (X0 = 0.01 g/L; Xmax = 2.948 g/L; μmax = 0.59/h; R2 = 0.97) was used for mathematical interpretation of biomass production. Mercier's model proved to be better than Luedeking–Piret model in describing β-galactosidase production (P0 = 0.7942 U/L; Pmax = 2169.3 U/L; Pr = 0.696/h; R2 = 0.99) whereas the latter was more efficient in mathematical illustration of lactose utilization (m = 0.187 g/g h; Yx/s = 0.301 g/L; R2 = 0.98) among the two used in this study. Strategies like fed-batch fermentation (3694.6 U/L) and semi-continuous fermentation (5551.9 U/L) further enhanced β-galactosidase production by 1.8 and 2.8 fold respectively.  相似文献   

6.
Modelling and simulation was done for a two-stage membrane-integrated hybrid reactor system for continuous production of L (+) lactic acid under non-neutralizing conditions. The model captures microbial conversion of sugar cane juice to lactic acid under substrate–product inhibitions with downstream purification by nanofiltration. All the major phenomena and the governing parameters like fluid flow, feed dilution, substrate–product inhibitions, Donnan and steric effects during micro and nanofiltration for cell recycle, product separation and purification have been reflected in the modelling. The model describes a green, integrated continuous process of direct lactic acid production starting with a cheap, renewable carbon source. The highest lactic acid concentration achieved after the final stage of nanofiltration was 66.97 g/L at 13 kg/cm2 operating pressure when the overall productivity reached 12.40 g/(L h). The developed model could successfully predict production, purification and transport of lactic acid through two stage membrane modules. Performance of the model was very good as indicated in the high overall correlation coefficient (R2 > 0.980) and the low relative error (RE < 0.1).  相似文献   

7.
The aim of the present study is to develop an efficient and cost-effective method for α-arbutin production by using whole-cell of Xanthomonas maltophilia BT-112 as a biocatalyst. Hydroquinone (HQ), substrate for the bioconversion as glucosyl acceptor, was immobilized on H107 macroporous resin to reduce its toxic effect on the cells, and the optimal reaction conditions for α-arbutin synthesis were investigated. When 350 g/L H107 resin (254.5 mM HQ) and 20 g/L (4.2 U/g) of cells were shaken in 10 mL Na2HPO4–KH2PO4 buffer (50 mM, pH 6.5) containing 509 mM sucrose at 35 °C with 150 rpm for 48 h, the final yield of α-arbutin reached 65.9 g/L with a conversion yield of 95.2% based on the amount of HQ supplied. The α-arbutin production was 202% higher than that of the control (free HQ) and the cells maintained its full activity for almost six consecutive batch reactions, indicating a potential for reducing production costs. Additionally, the product was one-step isolated and identified as α-arbutin by 13C NMR and 1H NMR analysis. In conclusion, the combination of whole cells and immobilized hydroquinone (IMHQ) is a promising approach for economical and industrial-scale production of α-arbutin.  相似文献   

8.
Hydroquinone glycosides were produced by transglycosylation reactions catalyzed by cyclodextrin glucanotransferase (CGTase) from Thermoanaerobacter sp. (Toruzyme® 3.0L). The reactions were carried out in an aqueous system containing hydroquinone (HQ) and maltodextrin as acceptor and donor substrate molecules respectively. The conditions for the synthesis of hydroquinone glucoside (α-arbutin) were 9 mM hydroquinone, maltodextrin (5%, w/v) in 20 mM citrate phosphate buffer, pH 5.5 and 0.025 mg/ml toruzyme at 40 °C for 24 h. The transfer efficiency of hydroquinone glycosylation was 31.8% and 29.2% respectively, when α-cyclodextrin and maltodextrin were employed as donor substrates. The major glycoside product was identified as hydroquinone-1-O-α-d-glucopyranoside (α-arbutin) on the basis of mass spectrometric, nuclear magnetic resonance analysis and component analysis of its enzymatic hydrolysates. The highest molar yield of α-arbutin (21.2%) was obtained when α-cyclodextrin was used as the donor substrate. A two step enzymatic reaction system comprising of CGTase and amyloglucosidase helped to attain a molar yield of 30% for α-arbutin. At room temperature the solubility of α-arbutin in water was determined to be 12.8 g/100 ml which is approximately 1.8 fold higher than that of hydroquinone.  相似文献   

9.
Jerusalem artichoke extract or powder was used for astaxanthin production using Phaffia rhodozyma without acidic or enzymatic inulin hydrolysis. The culture medium containing Jerusalem artichoke as carbon source was optimized, and feeding strategies, including constant, exponential, pH-stat, and substrate feedback fed-batch fermentations, were also compared for enhancing the cell biomass and astaxanthin synthesis by P. rhodozyma. Substrate-feedback fed-batch fermentation resulted in the highest dry cell weight of 83.60 g/L, with a carotenoid concentration and yield of 982.50 mg/L and 13.30 mg/g, respectively, under optimized medium components using Jerusalem artichoke extract as carbon source in a 3-L stirred-tank bioreactor. Moreover, 482.50 mg/L of carotenoids and 253.10 mg/L of astaxanthin were obtained by continuous feeding of Jerusalem artichoke powder, which was used as carbon source. Astaxanthin essence with high DPPH-scavenging activity was obtained from the extracted astaxanthin, and the DPPH free radical scavenging rate of 40 ppm astaxanthin essence reached 76.29%. When stored at 4 °C, astaxanthin essence showed the highest stability, with a minimum k value of 0.0099 week−1 and maximum half-life (t1/2) value of 70 weeks.  相似文献   

10.
Aerobic granular sludge was cultivated in a glass sequencing batch reactor (SBR) with glucose synthetic wastewater. The spherical shaped granules were observed on 4th day with the mean diameter of 0.1 mm. With the increase of chemical oxygen demand (COD) concentration of the influent, aerobic granules grew matured, the size of which ranged from 1.2 to 1.9 mm. The aerobic granular sludge could sustain high organic loading rate (about 4.0 g COD L−1 d−1), with good settling ability (settling velocity 36 m/h) and high biomass concentration (MLSS 6.7 ±0.2 g/L). Experimental data indicated that the substrate utilization and biomass growth kinetics followed Monod's kinetics model approximately. The corresponding kinetic coefficients of maximum specific substrate utilization rate (k), half velocity coefficient (Ks), growth yield coefficient (Y) and decay coefficient (Kd) were 13.2 d−1, 275.8 mg/L, 0.183–0.250 mg MLSS/mg COD and 0.023–0.075 d−1, respectively, which made aerobic granules have short setup period, high rate of substrate utilization and little surplus sludge.  相似文献   

11.
《Process Biochemistry》2014,49(5):740-744
The effects of oxido-reduction potential (ORP) control on succinic acid production have been investigated in Escherichia coli LL016. In LL016, two CO2 fixation pathways were achieved and NAD+ supply was enhanced by co-expression of heterologous pyruvate carboxylase (PYC) and nicotinic acid phosphoribosyltransferase (NAPRTase). During anaerobic fermentation, cell growth and metabolite distribution were changed with redox potential levels in the range of −200 to −400 mV. From the results, the ORP level of −400 mV was preferable, which resulted in the high succinic acid concentration (28.6 g/L) and high succinic acid productivity (0.33 g/L/h). Meanwhile, the yield of succinic acid at the ORP level of −400 mV was 39% higher than that at the ORP level of −200 mV. In addition, a higher NADH/NAD+ ratio and increased enzyme activities were also achieved by regulating the culture to a more reductive environment, which further enhanced the succinic acid production.  相似文献   

12.
We examined glucose 6-phosphate dehydrogenase (G6PD) production by fed-batch cultivation, using a recombinant strain of Saccharomyces cerevisiae W303-181 overexpressing this enzyme. The cultivations were carried out in a 3 L fermenter at pH 5.7, 30 °C, 2.0 vvm aeration, 200 rpm agitation and an inoculum concentration of 1.0 g/L. The volume of the culture medium in the fed-batch process varied from 1.333 to 2.0 L, due to the addition of 15.0 g/L glucose solution during 5 h. Different feeding rates were studied (exponentially increasing and decreasing feeding rates), and the feeding profile was determined by values of the parameter K (time constant), namely: 0.2, 0.5 and 0.8 h−1. The best enzyme production (847 U/L) was obtained with an exponentially increasing feeding rate and K = 0.2 h−1. The results attained also showed that this process is promising for G6PD production.  相似文献   

13.
Diacetyl, a highly valuable product that is extensively used as an ingredient of food, tobacco, and daily chemicals such as perfumes, can be produced from the nonenzymatic oxidative decarboxylation of α-acetolactate during bacterial fermentation and converted to acetoin and 2,3-butanediol by 2,3-butanediol dehydrogenase. In the present study, Bacillus sp. DL01, which gives high acetoin production, was metabolically engineered to improve diacetyl production. After the deletion of α-acetolactate decarboxylase (ALDC)-encoding gene (alsD) by homologous recombination, the engineered strain, named Bacillus sp. DL01-ΔalsD, lost ALDC activity and produced 1.53 g/L diacetyl without acetoin and 2,3-butanediol accumulation. The channeling of carbon flux into diacetyl biosynthetic pathway was amplified by an overexpressed α-acetolactate synthase (ALS)-encoding gene (alsS) in Bacillus sp. DL01-ΔalsD-alsS, which produced 4.02 g/L α-acetolactate and 1.94 g/L diacetyl, and the conversion from α-acetolactate to diacetyl was increased by 1-fold after 20 mM Fe3+ was added to the fermentation medium. A titer of 8.69 g/L diacetyl, the highest reported diacetyl production, was achieved by fed-batch fermentation in optimal conditions using the metabolically engineered strain of Bacillus sp. DL01-ΔalsD-alsS. These results are of great importance as a new method for the efficient production of diacetyl by food-safe bacteria.  相似文献   

14.
For this study, 2,3-butanediol (BD) fermentation from pure and biomass-derived sugar were optimized in shake-flask and 5-L bioreactor levels using Klebsiella oxytoca ATCC 8724. The results showed that 70 g/L of single sugar (glucose or xylose) and 90 g/L of mixed-sugar (glucose:xylose = 2:1) were optimum concentrations for efficient 2,3-BD fermentation. At optimum sugar concentrations, 2,3-BD productivities were 1.03, 0.64 and 0.50 gL−1 h−1, and yields were 0.43, 0.36 and 0.35 g/g in glucose, xylose and mixed-sugar medium, respectively. The lack of simultaneous utilization of glucose and xylose led to the lowest productivity in the mixed-sugar medium. Detoxification of biomass hydrolyzates was necessary for efficient 2,3-BD fermentation when sugar concentrations in the medium was 90 g/L or higher, but not with sugar concentrations of 30 g/L or less. A fed-batch fermentation using glucose medium led to an increase 2,3-BD titer to 79.4 g/L and yields 0.47 g/g, while productivity decreased to 0.79 gL−1 h−1. However, the fed-batch process was inefficient using mixed-sugar and biomass hydrolyzates because of poor xylose utilization. These results indicated that appropriate biomass processing technologies must be developed to generate separate glucose and xylose streams to produce high 2,3-BD titer from biomass-derived sugar using a fed-batch process.  相似文献   

15.
Streptomyces sp. Z2 was isolated from nitrobenzene contaminated activated sludge, which utilized nitrobenzene as a sole source of carbon, nitrogen, and energy under aerobic condition. It was found that besides nitrobenzene strain Z2 can degrade 2-picolinic acid. Strain Z2 completely degraded 2-picolinic acid with initial concentration of 500 mg/L, 1000 mg/L, 1500 mg/L, 2000 mg/L, 2500 mg/L, and 3000 mg/L within 36 h, 50 h, 72 h, 100 h, 136 h, and 180 h, respectively. Kinetics of 2-picolinic acid degradation was described using the Andrews equation. The kinetic parameters were as follows: qmax = 3.81 h?1, Ks = 83.10 mg/L, and Ki = 252.11 mg/L. During the biodegradation process, Z2 transformed 2-picolinic acid into a product which was identified as 6-hydroxy picolinic acid by UV–vis spectrometry, 1H nuclear magnetic resonance spectroscopy, and mass spectrometry. 6-Hydroxy picolinic acid was then cleaved and mineralized with release of ammonia.  相似文献   

16.
17.
2-Keto-l-gulonic acid (2-KLG), the direct precursor of vitamin C, is currently produced by a two-step fermentation route from d-sorbitol. However, this route involves three bacteria, making the mix-culture system complicated and redundant. Thus, replacement of the conventional two-step fermentation process with a one-step process could be revolutionary in vitamin C industry. In this study, different combinations of five l-sorbose dehydrogenases (SDH) and two l-sorbosone dehydrogenases (SNDH) from Ketogulonicigenium vulgare WSH-001 were introduced into Gluconobacter oxydans WSH-003, an industrial strain used for the conversion of d-sorbitol to l-sorbose. The optimum combination produced 4.9 g/L of 2-KLG. In addition, 10 different linker peptides were used for the fusion expression of SDH and SNDH in G. oxydans. The best recombinant strain (G. oxydans/pGUC-k0203-GS-k0095) produced 32.4 g/L of 2-KLG after 168 h. Furthermore, biosynthesis of pyrroloquinoline quinine (PQQ), a cofactor of those dehydrogenases, was enhanced to improve 2-KLG production. With the stepwise metabolic engineering of G. oxydans, the final 2-KLG production was improved to 39.2 g/L, which was 8.0-fold higher than that obtained using independent expression of the dehydrogenases. These results bring us closer to the final one-step industrial-scale production of vitamin C.  相似文献   

18.
The exponential substrate feeding (open-loop) and automated feedback substrate feeding (closed loop) strategies were developed to obtain high cell densities of fluorescent pseudomonad strains R62 and R81 and enhanced production of antifungal compound 2,4-diacetylphloroglucinol (DAPG) from glycerol as a sole carbon source. The exponential feeding strategy resulted in increased glycerol accumulation during the fed-batch cultivation when the predetermined specific growth rate (μ) was set at 0.10 or 0.20 h?1 (<μm = 0.29 h?1). Automated feeding strategies using dissolved oxygen (DO) or pH as feedback signals resulted in minimal to zero accumulation of glycerol for both the strains. In case of DO-based feeding strategy, biomass productivity of 0.24 g/(L h) and 0.29 g/(L h) was obtained for R62 and R81, respectively. Using pH-based feeding strategy, biomass productivity could be increased to a maximum of 0.51 and 0.54 g/(L h), for the strains R62 and R81, respectively, whereas the DAPG concentration was enhanced to 298 mg/L for R62 and 342 mg/L for R81 strains. These yields of DAPG are thus far the highest reported from GRAS organisms.  相似文献   

19.
Cellulosimicrobium cellulans employs extracellular sialidase to selectively convert polysialogangliosides to ganglioside GM1. We cloned this novel sialidase gene (ccsia) from C. cellulans sp. 21, and overexpressed recombinant sialidase (CcSia) protein in E. coli BL21 (DE3) by high cell density fermentation. The presence of an N-terminal hexa-His tag allowed for purification using nickel affinity chromatography (2.3-fold, specific activity 41.5 U/mg). As determined by gel electrophoresis and gel filtration chromatography, the molecular weight of CcSia was found to be about 75 kDa, consistent with sequence analysis (75,271 Da). CcSia transformed polysialogangliosides GD1a, GD1b and GT1b into GM1. For this reaction, the response surface approach showed that optimal conditions in a 1-L system were 2 h incubation at 32.5 °C and pH 5.2, with substrate concentrations of 10 g/L and crude enzyme concentration 1 g/L, respectively. Under above conditions, 10 g/L of ganglioside was completely converted to the product GM1 with a yield of 52%. Our studies demonstrate CcSia could be used for industrial preparation of ganglioside GM1 by the pharmaceutical industry.  相似文献   

20.
《Process Biochemistry》2014,49(12):2174-2180
Different filamentous fungi isolated from molasses and jams (kiwi and fig) were screened for fructooligosaccharides (FOS) producing activity. Two strains, identified as Penicilium sizovae (CK1) and Cladosporium cladosporioides (CF215), were selected on the basis of the FOS yield and kestose/nystose ratio. In both strains the activity was mostly mycelium-bound. Starting from 600 g/L of sucrose, maximum FOS yield was 184 and 339 g/L for P. sizovae and C. cladosporioides, respectively. Interestingly, the highest FOS concentration with C. cladosporioides was reached at 93% sucrose conversion, which indicated a notable transglycosylation to hydrolysis ratio. The main FOS in the reaction mixtures were identified by HPAEC–PAD chromatography. C. cladosporioides synthesized mainly 1-kestose (158 g/L), nystose (97 g/L), 1F-fructosylnystose (19 g/L), 6-kestose (12 g/L), neokestose (10 g/L) and a disaccharide (34 g/L) that after its purification and NMR analysis was identified as blastose [Fru-β(2  6)-Glc]. P. sizovae was very selective for the formation of 1F-FOS (in particular 1-kestose) with minor contribution of neoFOS and negligible of levan-type FOS.  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