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 390 毫秒
1.
The influence of oxygen transfer rate (OTR) on the molecular mass of alginate was studied. In batch cultures without dissolved oxygen tension (DOT) control and at different agitation rates, the DOT was nearly zero and the OTR was constant during biomass growth, hence the cultures were oxygen-limited. The OTR reached different maximum levels (OTRmax) and enabled to establish various relative respiration rates. Overall, the findings showed that OTR influences alginate molecular mass. The mean molecular mass (MMM) of the alginate increased as OTRmax decreased. The molecular mass obtained at 3.0 mmol l−1 h−1 was 7.0 times higher (1,560 kDa) than at 9.0 mmol l−1 h−1 (220 kDa). An increase in molecular mass can be a bacterial response to adverse nutritional conditions such as oxygen limitation.  相似文献   

2.
Oxygen supply is crucial in industrial application of microbial systems, such as Corynebacterium glutamicum, but oxygen transfer is often neglected in early strain characterizations, typically done under aerobic conditions. In this work, a new procedure for oxygen transfer screening is presented, assessing the impact of maximum oxygen transfer conditions (OTRmax) within microtiter plate-based cultivation for enhanced throughput. Oxygen-dependent growth and productivity were characterized for C. glutamicum ATCC13032 and C. glutamicum DM1933 (lysine producer). Biomass and lysine product yield are affected at OTRmax below 14 mmol L?1 h?1 in a standardized batch process, but not by further increase of OTRmax above this threshold value indicating a reasonable tradeoff between power input and oxygen transfer capacity OTRmax. The described oxygen transfer screening allows comparative determination of metabolic robustness against oxygen transfer limitation and serves identification of potential problems or opportunities later created during scale-up.  相似文献   

3.

Background  

The oxygen transfer rate (OTR) and dissolved oxygen tension (DOT) play an important role in determining alginate production and its composition; however, no systematic study has been reported about the independent influence of the OTR and DOT. In this paper, we report a study about alginate production and the evolution of the molecular mass of the polymer produced by a wild-type A. vinelandii strain ATCC 9046, in terms of the maximum oxygen transfer rate (OTRmax) in cultures where the dissolved oxygen tension (DOT) was kept constant.  相似文献   

4.
Alginate production and gene expression of genes involved in alginate biosynthesis were evaluated in continuous cultures under dissolved oxygen tension (DOT) controlled conditions. Chemostat at 8% DOT showed an increase in the specific oxygen uptake rate \((q_{{{\text{O}}_{ 2} }} )\) from 10.9 to 45.3 mmol g?1 h?1 by changes in the dilution rate (D) from 0.06 to 0.10 h?1, whereas under 1% DOT the \(q_{{{\text{O}}_{ 2} }}\) was not affected. Alginate molecular weight was not affected by DOT. However, chemostat at 1% DOT showed a downregulation up to 20-fold in genes encoding both the alginate polymerase (alg8, alg44), alginate acetylases (algV, algI) and alginate lyase AlgL. alyA1 and algE7 lyases gene expressions presented an opposite behavior by changing the DOT, suggesting that A. vinelandii can use specific depolymerases depending on the oxygen level. Overall, the DOT level have a differential effect on genes involved in alginate synthesis, thus a gene expression equilibrium determines the production of alginates of similar molecular weight under DOT controlled.  相似文献   

5.
Optimal agitation and aeration conditions [assuring O2 transfer rates (OTR) from 12 to 179 mmol L?1 h?1] were determined for pectin lyase (PL) synthesis of an Aspergillus niger strain. Components of the pectolytic enzyme complex were also investigated in order to determine whether their O2 demand is identical with or different from that of pectin lyase. Should the latter be the case, a possibility would be given to produce enzyme complexes of different agitation and aeration conditions. According to our results, mycelium yield of Aspergillus niger attained a maximum at an OTR of 100 mmol L?1 h?1. The yields of the various pectolytic enzymes reached maxima at different OTRs. Pectin lyase production was the highest (0.555 µmol min?1 mL?1) at an OTR of 60 mmol L?1 h?1. Endopolygalacturonase (PG) production showed a maximum at the OTR of 49 mmol L?1 h?1 with a second peak at 100?135 mmol O2 L?1 h?1. Pectin esterase (PE) synthesis showed a maximum at on OTR of 12?14 mmol L?1 h?1, while both apple juice clarifying and macerating activities gave two maxima at 14 and 60 mmol L?1 h?1 due to the optima of PE and endo-PG. Macerating activity showed a high value at OTR optimal for PL production as well.  相似文献   

6.
Sugar transport is very critical in developing an efficient and rapid conversion process of a mixture of sugars by engineered microorganisms. By using expressed sequence tag data generated for the fructophilic yeast Candida magnoliae JH110, we identified two fructose-specific transporters, CmFSY1 and CmFFZ1, which show high homology with known fructose transporters of other yeasts. The CmFSY1 and CmFFZ1 genes harbor no introns and encode proteins of 574 and 582 amino acids, respectively. Heterologous expression of the two fructose-specific transporter genes in a Saccharomyces cerevisiae, which is unable to utilize hexoses, revealed that both transporters are functionally expressed and specifically transport fructose. These results were further corroborated by kinetic analysis of the fructose transport that showed that CmFsy1p is a high-affinity fructose–proton symporter with low capacity (K M?=?0.13?±?0.01 mM, V max?=?2.1?±?0.3 mmol h?1 [gdw]?1) and that CmFfz1p is a low-affinity fructose-specific facilitator with high capacity (K M?=?105?±?12 mM, V max?=?8.6?±?0.7 mmol h?1 [gdw]?1). These fructose-specific transporters can be used for improving fructose transport in engineered microorganisms for the production of biofuels and chemicals from fructose-containing feedstock.  相似文献   

7.
Repeated-batch fermentation by a flocculating fusant, Saccharomyces cerevisiae HA 2, was done in a molasses medium that contained 20% (w/v) total sugar, at 30°C in an automatically controlled fermentor, and the effects of ethanol concentration on the specific growth rate and the specific production rate of ethanol were studied. Both the specific growth rate and the specific production rate of ethanol fell with increase of ethanol concentration, and there was a linear correlation between each rate and the concentration of thanol. The maximum specific growth rate (μmax) and the maximum specific production rate of ethanol (qmax) were 0.12 h−1 and 0.1 g ethanol/109 cells·h, respectively. The specific growth rate and the specific production rate of ethanol fell to zero at ethanol concentration of 89 g/l and 95 g/l, respectively. The number of viable cells, calculated from the linear inhibition equation, was 1.3 × 109 cells/ml for production of 85 g/l ethanol at a dilution rate (D1) of 0.2 h−1. Based on this estimation, a laboratory-scale continuous fermentation, using two fermentors in series, was done. In the second fermentor, 85 g/l ethanol was produced at a dilution rate (D1) of 0.2 h−1 by the active feedig of the fermented mash from the first fermentor into the second fermentor by pumping (hereafter called active feeding). To maintain the number of viable cells above 109 cells/ml in the second fermentor, a active feeding ratio of more than 23% was required. Under these conditions, 81 g/l ethanol was produced in the second fermentor at a dilution rate (Dt) of 0.25 h−1, and the high ethanol productivity of 20.3 g/l·h could be achieved. A bench-scale continuous fermentation, using two fermentors in series, with a active feeding ratio of 25% was done. An ethanol concentration of 84 g/l in the second fermentor at a dilution rate (Dt) of 0.25 h−1 was achieved, just as it was in the laboratory-scale fermentation test.  相似文献   

8.
Fermentations of Xanthomonas campestris, NRRL B-1459, were carried out in a bubble column fermentor (BCF) and in a stirred tank fermentor (STF) to allow comparison of representative variables measured during the microbial growth and the gum production. The microbial growth phase was described by a logistic rate equation where maximum cell concentration was provided by nitrogenous compounds balance. The average value of the maximum specific growth rate was higher in the bubble column (μ M =0.5 h?1) than in the stirred reactor (μ M =0.4 h?1). The upper values of xanthan yield (Y g-x =0.65 kg xanthan/kg glucose; Y O 2?x xanthan/kg oxygen) and specific production rate (q x =0.26 kg xanthan/kg biomass · h) were measured when the oxygen transfer coefficient was kept up above 80 h?1 in the STF fermentor. In the bubble column the fermentation achieved in the same culture medium lasts two times longer than in the stirred aerated tank; this was attributed to the low value of the oxygen transfer coefficient (K L a =20 h?1) at the beginning of the gum synthesis phase. The results obtained in the stirred tank were the basis to estimate the optimal biomass concentration which enables to achieve a culture in non-limiting oxygen transfer conditions. Nevertheless, the transfer characteristics were more homogeneous in the bubble column than in the stirred tank where dead stagnant zones were observed. This is of primary importance when establishing fermentation kinetics models.  相似文献   

9.
The impact of mass transfer on productivity can become a crucial aspect in the fermentative production of bulk chemicals. For highly aerobic bioprocesses the oxygen transfer rate (OTR) and productivity are coupled. The achievable space time yields can often be correlated to the mass transfer performance of the respective bioreactor. The oxygen mass transfer capability of a jet aerated loop reactor is discussed in terms of the volumetric oxygen mass transfer coefficient kLa [h?1] and the energetic oxygen transfer efficiency E [kgO2 kW?1 h?1]. The jet aerated loop reactor (JLR) is compared to the frequently deployed aerated stirred tank reactor. In jet aerated reactors high local power densities in the mixing zone allow higher mass transfer rates, compared to aerated stirred tank reactors. When both reactors are operated at identical volumetric power input and aeration rates, local kLa values up to 1.5 times higher are possible with the JLR. High dispersion efficiencies in the JLR can be maintained even if the nozzle is supplied with pressurized gas. For increased oxygen demands (above 120 mmol L?1 h?1) improved energetic oxygen transfer efficiencies of up to 100 % were found for a JLR compared to an aerated stirred tank reactor operating with Rushton turbines.  相似文献   

10.
To grow yeast rapidly, it is necessary to supply sufficient oxygen to the yeast and to effectively remove the heat of the fermentation. We succeeded in developing a commercial-scale fermentor for growing a food yeast (Candida utilis) to produce RNA rapidly. This fermentor is an internal-loop airlift type with vertical heat transfer tubes between inner and the outer columns. The volume of the fermentor is 145 m3 (working volume 75 m3). The oxygen transfer rate (OTR) was 9.9 kg-O2/m3/h using a superficial gas velocity of 30 cm/s based on the outer column. Much of the heat of fermentation and the energy resulting from aeration could be removed effectively by the heat transfer tubes. This unique airlift fermentor was driven at a dilution rate of 0.43 h−1 for about 70 d, with the yeast concentration being maintained at 22.8 kg-dry cell/m3. The yeast production rate was 9.79 kg-dry cell/m3/h. Compared with a traditional stirred-type fermentor, two Vogelbush-type fermentors and another airlift fermentor, our fermentor was far superior with respect to OTR and yeast productivity.  相似文献   

11.
High Pichia pastoris biomass density could be obtained using high co-feeding rate of methanol and sorbitol in a fed-batch or continuous culture, while further higher feeding rate finally leads to oxygen limitation in bioreactor. In the literature, there is lack of report about AOX1 promoter regulation with regard to dissolved oxygen level (DO). Therefore, in this work, chemostat cultures were performed to investigate the cell growth, metabolism and regulation of the AOX1 promoter (pAOX1) regarding co-feeding rate of optimized methanol/sorbitol mixture (methanol fraction 0.60 C-mol/C-mol) using a P. pastoris Mut+/pAOX1-lacZ strain. The oxygen transfer rates (OTR) in bioreactor were kept in the range of typical values of large bioreactor, i.e., 4–8 g/(L h) if DO equals 30 % saturation or 5–10 g/(L h) if DO nears zero. For DO >0, an increase of the carbon fed led to an increase of pAOX1 induction. By contrast, when dissolved oxygen was completely depleted, methanol accumulated, causing a 30 % decrease of pAOX1 induction. However, this decrease is more likely to be lined to methanol accumulation than to low level of dissolved oxygen (<4 % DO). Methanol/sorbitol co-feeding allowed cells to adapt to oxygen transient limitations that often occur at industrial scale with reduced effect on pAOX1 induction. The optimal feeding rate tested here was 6.6 mmol C (DCW h)?1 at an OTR of 8.28 g O2(L h)?1 with over fivefold pAOX1 induction (probably directly associated with target protein productivity) compared with previous work.  相似文献   

12.
The aim of the present work was to study the effects of photoperiod, salinity and pH on growth and lipid content of Pavlova lutheri microalgae for biodiesel production in small-scale and large-scale open-pond tanks. In a 250-mL flask, the cultures grew well under 24 h illumination with maximum specific growth rate, μ max , of 0.12 day?1 and lipid content of 35 % as compared to 0.1 day?1 and 15 % lipid content in the dark. The salinity was optimum for the cell growth at 30–35 ppt, but the lipid content of 34–36 % was higher at 35–40 ppt. Algal growth and lipid accumulation was optimum at pH 8–9. Large-scale cultivation in 5-L and 30-L tanks achieved μ max of 0.13–0.14 day?1 as compared to 0.12 day?1 in small-scale and 300L cultures.  相似文献   

13.
Oxygen-responsive promoters can be useful for synthetic biology applications, however, information on their characteristics is still limited. Here, we characterized a group of heterologous microaerobic globin promoters in Escherichia coli. Globin promoters from Bacillus subtilis, Campylobacter jejuni, Deinococcus radiodurans, Streptomyces coelicolor, Salmonella typhi and Vitreoscilla stercoraria were used to express the FMN-binding fluorescent protein (FbFP), which is a non-oxygen dependent marker. FbFP fluorescence was monitored online in cultures at maximum oxygen transfer capacities (OTRmax) of 7 and 11 mmol L?1 h?1. Different FbFP fluorescence intensities were observed and the OTRmax affected the induction level and specific fluorescence emission rate (the product of the specific fluorescence intensity multiplied by the specific growth rate) of all promoters. The promoter from S. typhi displayed the highest fluorescence emission yields (the quotient of the fluorescence intensity divided by the scattered light intensity at every time-point) and rate, and together with the promoters from D. radiodurans and S. coelicolor, the highest induction ratios. These results show the potential of diverse heterologous globin promoters for oxygen-limited processes using E. coli.  相似文献   

14.
The hybridoma 192 was used to produce a monoclonal antibody (MAb) against 17‐hydroxyprogesterone (17‐OHP), for possible use in screening for congenital adrenal hyperplasia (CAH). The factors influencing the MAb production were screened and optimized in a 2 L stirred bioreactor. The production was then scaled up to a 20 L bioreactor. All of the screened factors (aeration rate, stirring speed, dissolved oxygen concentration, pH, and temperature) were found to significantly affect production. Optimization using the response surface methodology identified the following optimal production conditions: 36.8°C, pH 7.4, stirring speed of 100 rpm, 30% dissolved oxygen concentration, and an aeration rate of 0.09 vvm. Under these conditions, the maximum viable cell density achieved was 1.34 ± 0.21 × 106 cells mL?1 and the specific growth rate was 0.036 ± 0.004 h?1. The maximum MAb titer was 11.94 ± 4.81 μg mL?1 with an average specific MAb production rate of 0.273 ± 0.135 pg cell?1 h?1. A constant impeller tip speed criterion was used for the scale‐up. The specific growth rate (0.040 h?1) and the maximum viable cell density (1.89 × 106 cells mL?1) at the larger scale were better than the values achieved at the small scale, but the MAb titer in the 20 L bioreactor was 18% lower than in the smaller bioreactor. A change in the culture environment from the static conditions of a T‐flask to the stirred bioreactor culture did not affect the specificity of the MAb toward its antigen (17‐OHP) and did not compromise the structural integrity of the MAb. © 2012 American Institute of Chemical Engineers Biotechnol. Prog., 2013  相似文献   

15.
Production of lipopeptides fengycin and surfactin in rotating discs bioreactor was studied. The effects of rotation velocity and the addition of agitators between the discs on volumetric oxygen transfer coefficient k L a were firstly studied in model media. Then the production of lipopeptides was also studied at different agitation conditions in the modified bioreactor (with agitators). The effect of agitation on dissolved oxygen, on submerged and immobilized biomass, on lipopeptide concentrations and yields and on the selectivity of the bioreaction was elucidated and discussed. The proposed modified rotating discs bioreactor allowed to obtain high fengycin concentrations (up to 787 mg L?1), but also better selectivity of the bioreaction towards fengycin (up to 88 %) and better yields of fengycin per glucose (up to 62.9 mg g?1), lipopeptides per glucose (up to 71.5 mg g?1), fengycin per biomass (up to 309 mg g?1) and lipopeptides per biomass (up to 396 mg g?1) than those reported in the literature. Highest fengycin production and selectivity were obtained at agitation velocity of 30 min?1. The proposed non-foaming fermentation process could contribute to the scale-up of lipopeptide fermentors and promote the industrial production of fengycin. The proposed bioreactor and bioprocess could be very useful also for the production of other molecules using bioprocesses requiring bubbleless oxygen supply.  相似文献   

16.
Alginate production by Azotobacter vinelandii growing in chemostat cultures was evaluated under different O2 transfer rates (OTR). As a result of modifying the culture’s agitation rate from 300 to 500 rpm, the OTR increased from 9 to 15.1 mmol l−1 h−1 and a slight variation in the alginate production (1.7–2.2 g l−1) was observed. At a constant growth rate (0.1 h−1), the mean molecular mass of the alginate was strongly influenced by changes in the OTR, varying from 860 to 1,690 kDa. These results support a possible relationship between alginate polymerization-depolymerization process and the O2 uptake rate.  相似文献   

17.
In this study, a compressed oxygen gas supply was connected to a sealed aerated stirred tank reactor (COS-SSTR) bio-system, leading to a high-oxygen pressure bioreactor used to improve the bio-transformative performance in the production of 1,3-dihydroxyacetone (DHA) from glycerol using Gluconobacter oxydans NL71. A concentration of 301.2 ± 8.2 g L?1 DHA was obtained from glycerol after 32 h of fed-batch fermentation in the COS-SSTR system. The volumetric productivity for this process was 9.41 ± 0.23 g L?1 h?1, which is presently the highest obtained level of glycerol bioconversion into DHA. These results show that the application of this bioreactor would enable microbial production of DHA from glycerol at the industrial scale.  相似文献   

18.
Biodegradation of pyridine by a novel bacterial strain, Rhizobium sp. NJUST18, was studied in batch experiments over a wide concentration range (from 100 to 1,000 mg l?1). Pyridine inhibited both growth of Rhizobium sp. NJUST18 and biodegradation of pyridine. The Haldane model could be fitted to the growth kinetics data well with the kinetic constants μ* = 0.1473 h?1, K s = 793.97 mg l?1, K i = 268.60 mg l?1 and S m = 461.80 mg l?1. The true μ max, calculated from μ*, was found to be 0.0332 h?1. Yield coefficient Y X/S depended on S i and reached a maximum of 0.51 g g?1 at S i of 600 mg l?1. V max was calculated by fitting the pyridine consumption data with the Gompertz model. V max increased with initial pyridine concentration up to 14.809 mg l?1 h?1. The q S values, calculated from $V_{ \hbox{max} }$ , were fitted with the Haldane equation, yielding q Smax = 0.1212 g g?1 h?1 and q* = 0.3874 g g?1 h?1 at S m′ = 507.83 mg l?1, K s′ = 558.03 mg l?1, and K i′ = 462.15 mg l?1. Inhibition constants for growth and degradation rate value were in the same range. Compared with other pyridine degraders, μ max and S m obtained for Rhizobium sp. NJUST18 were relatively high. High K i and K i′ values and extremely high K s and K s′ values indicated that NJUST18 was able to grow on pyridine within a wide concentration range, especially at relatively high concentrations.  相似文献   

19.
N-Nitrosodimethylamine (NDMA) is an emerging contaminant of concern. N-nitrodimethylamine (DMNA) is a structural analog to NDMA. NDMA and DMNA have been found in drinking water, groundwater, and other media and are of concern due their toxicity. The authors evaluated biotransformation of NDMA and DMNA by cultures enriched from contaminated groundwater growing on benzene, butane, methane, propane, or toluene. Maximum specific growth rates of enriched cultures on butane (μmax = 1.1 h?1) and propane (μmax = 0.65 h?1) were 1 to 2 orders of magnitude higher than those presented in the literature. Growth rates of mixed cultures grown on benzene (μmax = 1.3 h?1), methane (μmax = 0.09 h?1), and toluene (μmax = 0.99 h?1) in these studies were similar to those presented in the literature. NDMA biotransformation rates for methane oxidizers (υmax = 1.4 ng min?1 mg?1) and toluene oxidizers (υmax = 2.3 ng min?1 mg?1) were comparable to those presented in the literature, whereas the biotransformation rate for propane oxidizers (υmax = 0.37 ng min?1 mg?1) was lower. NDMA biotransformation rates for benzene oxidizers (υmax = 1.02 ng min?1 mg?1) and butane oxidizers (υmax = 1.2 ng min?1 mg?1) were comparable to those reported for other primary substrates. These studies showed that DMNA biotransformation rates for benzene (υmax = 0.79 ng min?1 mg?1), butane (υmax = 1.0 ng min?1 mg?1), methane (υmax = 2.1 ng min?1 mg?1), propane (υmax = 1.46 ng min?1 mg?1), and toluene (υmax = 0.52 ng min?1 mg?1) oxidizers were all comparable. These studies highlight potential bioremediation methods for NDMA and DMNA in contaminated groundwater.  相似文献   

20.
The kinetics of chromium(VI) reduction by Pseudomonas aeruginosa (P. aeruginosa) and Escherichia coli (E. coli) was studied under both pure and mixed cultures. Initially, the study of kinetics was performed in pure culture. It was observed that the growth of the two bacteria was both inhibited in the presence of chromium(VI). The maximum specific growth rate (μ m ) of P. aeruginosa decreased from 2.3942 h?1 (without Cr(VI)) to 1.8551 h?1 (with Cr(VI)). Under the mixed culture, the growth of E. coli was inhibited by P. aeruginosa. The maximum specific growth rate (μ m ) of E. coli decreased form 0.871 h?1 (in pure culture) to 0.153 h?1 (in mixed culture). When the concentration of each bacterium was 4.5 × 108 cells ml?1, the half-velocity reduction rate constant (K C) and the maximum specific reduction rate constant (v max) of chromium(VI) were 80.05 mg chromium(VI) l?1 and 3.674 mg chromium(VI) cells?1 h?1, respectively. The results showed that the simulation appeared in good agreement with the experimental data, supporting the series of mathematical models represented the bacteria growth and chromium(VI) reduction in both pure and mixed cultures usefully.  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