首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
To check for possible mass transfer limitations of oxygen and/or carbon dioxide in kinetic experiments on microbial desulphurization of coal, it is important to properly measure the volumetric mass transfer coefficient (k(L)a) especially at high slurry densities. Volumetric mass transfer coefficients of oxygen, at different solid hold-up values (epsilon(s) = 0 to 0.28) of coal slurries (d(par) < 100 * 10(-6) m), were measured in a lab scale fermentor and in a lab scale pachuca tank, using the dynamic gas-liquid absorption method. It was shown that serious errors could occur due to oxygen adsorption at the coal surface. Using the data of an independently measured adsorption isotherm, the real k(L)a could be calculated from the measured apparent k(L)a. The results show a k(L)a decrease of 40% to 50% at a volumetric solid hold-up of 28%. Estimation of the oxygen and carbon dioxide transfer rates, from the measured mass transfer coefficients, indicates that the stirred fermentor is suitable for kinetic experiments at high slurry densities, whereas the pachuca tank and shake flask are not. (c) 1992 John Wiley & Sons, Inc.  相似文献   

2.
In order to see the effect of CO(2) inhibition resulting from the use of pure oxygen, we carried out a comparative fed-batch culture study of polyhydroxybutyric acid (PHB) production by Ralstonia eutropha using air and pure oxygen in 5-L, 30-L, and 300-L fermentors. The final PHB concentrations obtained with pure O(2) were 138.7 g/L in the 5-L fermentor and 131.3 g/L in the 30-L fermentor, which increased 2.9 and 6.2 times, respectively, as compared to those obtained with air. In the 300-L fermentor, the fed-batch culture with air yielded only 8.4 g/L PHB. However, the maximal CO(2) concentrations in the 5-L fermentor increased significantly from 4.1% (air) to 15.0% (pure O(2)), while it was only 1.6% in the 30-L fermentor with air, but reached 14.2% in the case of pure O(2). We used two different experimental methods for evaluating CO(2) inhibition: CO(2) pulse injection and autogenous CO(2) methods. A 10 or 22% (v/v) CO(2) pulse with a duration of 3 or 6 h was introduced in a pure-oxygen culture of R. eutropha to investigate how CO(2) affects the synthesis of biomass and PHB. CO(2) inhibited the cell growth and PHB synthesis significantly. The inhibitory effect became stronger with the increase of the CO(2) concentration and pulse duration. The new proposed autogenous CO(2) method makes it possible to place microbial cells under different CO(2) level environments by varying the gas flow rate. Introduction of O(2) gas at a low flow rate of 0.42 vvm resulted in an increase of CO(2) concentration to 30.2% in the exit gas. The final PHB of 97.2 g/L was obtained, which corresponded to 70% of the PHB production at 1.0 vvm O(2) flow rate. This new method measures the inhibitory effect of CO(2) produced autogenously by cells through the entire fermentation process and can avoid the overestimation of CO(2) inhibition without introducing artificial CO(2) into the fermentor.  相似文献   

3.
The rate constants and delta H degrees for the non-cooperative dimeric Busycon myoglobin are: oxygen, k' = 4.75 X 10(7) M-1 sec-1, k = 71 sec-1, and CO, l'= 3.46 X 10(5) M-1 sec-1, l = 0.0052 sec-1 at 20 degrees C, pH 7, delta H degrees = -3 kcal/mol for O2 and CO.2. Log-log plots of k vs K for oxygen and of l' vs L for CO binding for numerous non-cooperative hemoglobins and myoglobins point to a large steric influence of the protein on heme ligation reactions. Many of the proteins behave as "R" state for one ligand, but "T" for the other.  相似文献   

4.
Thermophilic (50 degrees C) and obligately thermophilic (60 degrees C) anaerobic carbohydrate- and protein-digesting and methanogenic bacterial populations were enumerated in a mesophilic (35 degrees C) fermentor anaerobically digesting municipal primary sludge. Of the total bacterial population in the mesophilic fementor, 9% were thermophiles (36 x 10(6)/ml) and 1% were obligate thermophiles (4.5 x 10(6)/ml). Of these 10%, the percentages of bacteria (thermophiles and obligate thermophiles, respectively) able to use specific substrates were further enumerated as follows: bacteria able to digest albumin, casein, starch, and mono- and disaccharides, 30 and 10%; pectin degraders, 10 and 0.2%; cellulose degraders, 2 and 0.06%; methanogens that grow with H2 and CO2, methanol, and dimethylamine, 9 and 1%; methanogens that grow with formate, 8 and 5%; and methanogens that grow with acetate, 25 and less than 0.8%. Shortly after the temperature was elevated from 35 to 50 or 60 degrees C, the digestion of albumin, casein, starch, and mono- and disaccharides was detected, and methane was produced from H2 and CO2. Methane produced from acetate was not delayed at 50 degrees C, but was delayed by 29 days at 60 degrees C. Methane produced from formate was delayed by 3 days, from methanol by 7 days, and from dimethylamine by 5 days at 50 and 60 degrees C. A 10- and 20-day acclimation period was required for hydrolysis of pectin and cellulose, respectively, at 50 degrees C. Digestion of pectin required 20 days and cellulose longer than 85 days when the temperature was elevated abruptly from 35 to 60 degrees C. The acclimation period for the digestion of pectin and cellulose at 60 degrees C was shortened to 3 and 15 days, respectively, by seeding with a small amount of a culture acclimated to 50 degrees C. The data suggest that enrichment of cellulolytic, pectinolytic, and acetate-utilizing bacteria is crucial for the digestion of sewage sludge at 60 degrees C.  相似文献   

5.
M Chen 《Applied microbiology》1983,45(4):1271-1276
Thermophilic (50 degrees C) and obligately thermophilic (60 degrees C) anaerobic carbohydrate- and protein-digesting and methanogenic bacterial populations were enumerated in a mesophilic (35 degrees C) fermentor anaerobically digesting municipal primary sludge. Of the total bacterial population in the mesophilic fementor, 9% were thermophiles (36 x 10(6)/ml) and 1% were obligate thermophiles (4.5 x 10(6)/ml). Of these 10%, the percentages of bacteria (thermophiles and obligate thermophiles, respectively) able to use specific substrates were further enumerated as follows: bacteria able to digest albumin, casein, starch, and mono- and disaccharides, 30 and 10%; pectin degraders, 10 and 0.2%; cellulose degraders, 2 and 0.06%; methanogens that grow with H2 and CO2, methanol, and dimethylamine, 9 and 1%; methanogens that grow with formate, 8 and 5%; and methanogens that grow with acetate, 25 and less than 0.8%. Shortly after the temperature was elevated from 35 to 50 or 60 degrees C, the digestion of albumin, casein, starch, and mono- and disaccharides was detected, and methane was produced from H2 and CO2. Methane produced from acetate was not delayed at 50 degrees C, but was delayed by 29 days at 60 degrees C. Methane produced from formate was delayed by 3 days, from methanol by 7 days, and from dimethylamine by 5 days at 50 and 60 degrees C. A 10- and 20-day acclimation period was required for hydrolysis of pectin and cellulose, respectively, at 50 degrees C. Digestion of pectin required 20 days and cellulose longer than 85 days when the temperature was elevated abruptly from 35 to 60 degrees C. The acclimation period for the digestion of pectin and cellulose at 60 degrees C was shortened to 3 and 15 days, respectively, by seeding with a small amount of a culture acclimated to 50 degrees C. The data suggest that enrichment of cellulolytic, pectinolytic, and acetate-utilizing bacteria is crucial for the digestion of sewage sludge at 60 degrees C.  相似文献   

6.
Taly A  Baciou L  Sebban P 《FEBS letters》2002,532(1-2):91-96
Photosynthetic reaction centers (RCs) from Rhodobacter sphaeroides were incorporated in dimyristoylphosphatidylcholine (DMPC) liposomes. The first and second electron transfer rates (k(AB)(1) and k(AB)(2), respectively) between the first and the second quinone electron acceptors have been measured as a function of temperature, across the phase transition of DMPC (23 degrees C). The Eyring plots of k(AB)(1) display straight lines. In contrast, the Eyring plots for k(AB)(2) in proteoliposomes show a break at about 23.5 degrees C. This physical discrimination between the two electron transfer reactions demonstrates that the stiffness of the lipid environment of the RCs and/or the protein-protein interactions influence the parameters governing k(AB)(2), but not the gating process limiting k(AB)(1).  相似文献   

7.
We describe a study of oxygen transfer in shake flasks using a non-invasive optical sensor. This study investigates the effect of different plugs, presence of baffles, and the type of media on the dissolved oxygen profiles during Escherichia coli fermentation. We measured the volumetric mass transfer coefficient (k(L)a) under various conditions and also the resistances of the various plugs. Finally, we compared shake flask k(L)a with that from a stirred tank fermentor. By matching k(L)a's we were able to obtain similar growth and recombinant protein product formation kinetics in both a fermentor and a shake flask. These results provide a quantitative comparison of fermentations in a shake flask vs. a bench-scale fermentor and should be valuable in guiding scale-up efforts.  相似文献   

8.
CO(2) transfer conductance from the intercellular airspaces of the leaf into the chloroplast, defined as mesophyll conductance (g(m)), is finite. Therefore, it will limit photosynthesis when CO(2) is not saturating, as in C3 leaves in the present atmosphere. Little is known about the processes that determine the magnitude of g(m). The process dominating g(m) is uncertain, though carbonic anhydrase, aquaporins, and the diffusivity of CO(2) in water have all been suggested. The response of g(m) to temperature (10 degrees C-40 degrees C) in mature leaves of tobacco (Nicotiana tabacum L. cv W38) was determined using measurements of leaf carbon dioxide and water vapor exchange, coupled with modulated chlorophyll fluorescence. These measurements revealed a temperature coefficient (Q(10)) of approximately 2.2 for g(m), suggesting control by a protein-facilitated process because the Q(10) for diffusion of CO(2) in water is about 1.25. Further, g(m) values are maximal at 35 degrees C to 37.5 degrees C, again suggesting a protein-facilitated process, but with a lower energy of deactivation than Rubisco. Using the temperature response of g(m) to calculate CO(2) at Rubisco, the kinetic parameters of Rubisco were calculated in vivo from 10 degrees C to 40 degrees C. Using these parameters, we determined the limitation imposed on photosynthesis by g(m). Despite an exponential rise with temperature, g(m) does not keep pace with increased capacity for CO(2) uptake at the site of Rubisco. The fraction of the total limitations to CO(2) uptake within the leaf attributable to g(m) rose from 0.10 at 10 degrees C to 0.22 at 40 degrees C. This shows that transfer of CO(2) from the intercellular air space to Rubisco is a very substantial limitation on photosynthesis, especially at high temperature.  相似文献   

9.
Specific dynamic action (SDA), the accumulated energy expended on all physiological processes associated with meal digestion, is strongly influenced by features of both the meal and the organism. We assessed the effects of meal size, meal type, body temperature, and body size on the postprandial metabolic response and calculated SDA of the marine toad, Bufo marinus. Peak postprandial rates of O(2) consumption (.V(O2)) and CO(2) production (.V(CO2)) and SDA increased with meal size (5%-20% of body mass). Postprandial metabolism was impacted by meal type; the digestion of hard-bodied superworms (Zophobas larva) and crickets was more costly than the digestion of soft-bodied earthworms and juvenile rats. An increase in body temperature (from 20 degrees to 35 degrees C) altered the postprandial metabolic profile, decreasing its duration and increasing its magnitude, but did not effect SDA, with the cost of meal digestion remaining constant across body temperatures. Allometric mass exponents were 0.69 for standard metabolic rate, 0.85 for peak postprandial .V(O2), and 1.02 for SDA; therefore, the factorial scope of peak postprandial .V(O2) increased with body mass. The mass of nutritive organs (stomach, liver, intestines, and kidneys) accounted for 38% and 20% of the variation in peak postprandial .V(O2) and SDA, respectively. Toads forced to exercise experienced 25-fold increases in .V(O2) much greater than the 5.5-fold increase experience during digestion. Controlling for meal size, meal type, and body temperature, the specific dynamic responses of B. marinus are similar to those of the congeneric Bufo alvarius, Bufo boreas, Bufo terrestris, and Bufo woodhouseii.  相似文献   

10.
Radiolabeling of biologically active molecules with the [(99m)Tc(CO)(3)](+) unit has been of primary interest in recent years. With this in mind, we herein report symmetric (L(1)) and asymmetric (L(2)-L(5)) pyrazolyl-containing chelators that have been evaluated in radiochemical reactions with the synthon [(99m)Tc(H(2)O)(3)(CO)(3)](+) (1a). These reactions yielded the radioactive building blocks [(99m)Tc(CO)(3)(k(3)-L)](+) (L = L(1)-L(5), 2a-6a), which were identified by RP-HPLC. The corresponding Re surrogates (2-6) allowed for macroscopic identification of the radiochemical conjugates. Complexes 2a-6a, with log P(o/w) values ranging from -2.35 to 0.87, were obtained in yields of > or =90% using ligand concentrations in the 10(-5-)10(-4) M range. Challenge studies with cysteine and histidine revealed high stability for all of these radioactive complexes, and biodistribution studies in mice indicated a fast rate of blood clearance and high rate of total radioactivity excretion, occurring primarily through the renal-urinary pathway. Based on the framework of the asymmetric chelators, the novel bifunctional ligands 3,5-Me(2)-pz(CH(2))(2)N((CH(2))(3)COOH)(CH(2))(2)NH(2) (L(6)) and pz(CH(2))(2)N((CH(2))(3)COOH)(CH(2))(2)NH(2) (L(7)) have been synthesized and their coordination chemistry toward (NEt(4))(2)[ReBr(3)(CO)(3)] (1) has been explored. The resulting complexes, fac-[Re(CO)(3)(k(3)-L)]Br (L(6)(7), L(7)(8)), contain tridentate ancillary ligands that are coordinated to the metal center through the pyrazolyl and amine nitrogen atoms, as observed for the other related building blocks. L(6) and L(7) were coupled to a glycylglycine ethyl ester dipeptide, and the resulting functionalized ligands were used to prepare the model complexes fac-[Re(CO)(3)(kappa(3)-3,5-Me(2)-pz(CH(2))(2)N(glygly)(CH(2))(2)NH(2))](+) (9/9a) and fac-[Re(CO)(3)(kappa(3)-pz(CH(2))(2)N(CH(2))(3)(glygly)(CH(2))(2)NH(2))](+) (10/10a) (M = Re, (99m)Tc). These small conjugates have been fully characterized and are reported herein. On the basis of the in vitro/in vivo behavior of the model complexes (2a-6a, 9a, 10a), we chose to evaluate the in vitro/in vivo biological behavior of a new tumor-seeking Bombesin pyrazolyl conjugate, [(L(6))-G-G-G-Q-W-A-V-G-H-L-M-NH(2)], that has been labeled with the [(99m)Tc(CO)(3)](+) metal fragment. Stability, in vitro cell binding assays, and pharmacokinetics studies in normal mice are reported herein.  相似文献   

11.
The stability of the N-terminal domain of the ribosomal protein L9, NTL9, from Bacillus stearothermophilus has been monitored by circular dichroism at various temperatures and chemical denaturant concentrations in H2O and D2O. The basic thermodynamic parameters for the unfolding reaction, deltaH(o), deltaS(o), and deltaC(o)p, were determined by global analysis of temperature and denaturant effects on stability. The data were well fit by a model that assumes stability varies linearly with denaturant concentration and that uses the Gibbs-Helmholtz equation to model changes in stability with temperature. The results obtained from the global analysis are consistent with information obtained from individual thermal and chemical denaturations. NTL9 has a maximum stability of 3.78 +/- 0.25 kcal mol(-1) at 14 degrees C. DeltaH(o)(25 degrees C) for protein unfolding equals 9.9 +/- 0.7 kcal mol(-1) and TdeltaS(o)++(25 degrees C) equals 6.2 +/- 0.6 kcal mol(-1). DeltaC(o)p equals 0.53 +/- 0.06 kcal mol(-1) deg(-1). There is a small increase in stability when D2O is substituted for H2O. Based on the results from global analysis, NTL9 is 1.06 +/- 0.60 kcal mol(-1) more stable in D2O at 25 degrees C and Tm is increased by 5.8 +/- 3.6 degrees C in D2O. Based on the results from individual denaturation experiments, NTL9 is 0.68 +/- 0.68 kcal mol(-1) more stable in D2O at 25 degrees C and Tm is increased by 3.5 +/- 2.1 degrees C in D2O. Within experimental error there are no changes in deltaH(o) (25 degrees C) when D2O is substituted for H2O.  相似文献   

12.
Fluctuations in pH and head-space pressure in a fermentor introduce temporary changes in off-gas CO(2) concentrations. These changes are quantified using a simple model based on kinetics of CO(2) hydration and gas-liquid mass transfer. The model is verified experimentally. An eigenvalue analysis of the model indicates that mass transfer is the parameter which controls the dynamics of CO(2) equilibration.  相似文献   

13.
The reaction between reduced Pseudomonas cytochrome c551 and cytochrome oxidase with two inorganic metal complexes, Co(phen)3(3+) and Mn(CyDTA)(H2O)-, has been followed by stopped-flow spectrophotometry. The electron transfer to cytochrome c551 by both reactants is a simple process, characterized by the following second-order rate constant: k = 4.8 X 10(4) M-1 sec-1 in the case of Co(phen)3(3+) and k = 2.3 X 10(4) M-1 sec-1 in the case of Mn(CyDTA)(H2O)-. The reaction of the c-heme of the oxidase with both metal complexes is somewhat heterogeneous, the overall process being characterized by the following second-order rate constants: k = 1.7 X 10(3) M-1 sec-1 with Co(phen)3(3+) and k = 4.3 X 10(4) M-1 sec-1 with Mn(CyDTA)(H2O)- as oxidants; under CO (which binds to the d1-heme) the former constant increases by a factor of 2, while the latter does not change significantly. The oxidation of the d1-heme of the oxidase by Co(phen)3(3+) occurs via intramolecular electron transfer to the c-heme, a direct bimolecular transfer from the complex being operative only at high metal complex concentrations; when Mn(CyDTA)(H2O)- is the oxidant, the bimolecular oxidation of the d1-heme competes successfully with the intramolecular electron transfer.  相似文献   

14.
The effects of rheological change by addition of sodium carboxymethylcellulose (CMC) to culture medium in an air-lift-type fermentor on autotrophic production of poly-(D-3-hydroxybutyric acid) [P(3HB)] by two-stage culture of Alcaligenes eutrophus is investigated. Addition of 0.05% CMC increased P(3HB) production rate during the P(3HB) accumulation phase to twice that of the control culture. It was thought that addition of a small amount of CMC was beneficial for production of P(3HB) employing the air-lift fermentor under safe autotrophic culture conditions in wich oxygen concentration was maintained below 6.9% (v/v). the volumetric mass transfer coefficient (K(L)a) observed in the presence of CMC is shown to correlated with the P(3HB) production rate obtained. (c) 1997 John Wiley & Sons, Inc. Biotechnol Bioeng 53: 529-533, 1997.  相似文献   

15.
? Respiratory acclimation to growth temperature differs between species, but underlying mechanisms are poorly understood. In the present study, we tested the hypothesis that respiratory acclimation of CO(2) release is a consequence of growth regulation such that growth rates of young foliage of Eucalyptus spp. are similar at contrasting growth temperatures. Further, we tested whether such a response is affected by adaptation of Eucalyptus to different thermal environments via growth at different altitudes in the Australian Alps. ? We employed calorimetric methods to relate rates of CO(2) release (mainly from substrate oxidation) and rates of O(2) reduction to conservation of energy. Temperature responses of these processes provided insight into mechanisms that control energy conservation and expenditure, and helped define 'instantaneous enthalpic growth capacity' (CapG). ? CapG increased with altitude, but was counteracted by other factors in species adapted to highland habitats. The acclimation response was partly driven by changes in respiratory capacity (CapR(CO2)), and partly by more pronounced dynamic responses of CO(2) release (δ(R(CO2))) to measurement temperature. We observed enhanced temperature sensitivity of O(2) reduction (E(o)(R(O2))) at higher altitudes. ? Adaptation to growth temperature included differences in respiration and growth capacities, but there was little evidence that Eucalyptus species vary in metabolic flexibility.  相似文献   

16.
Pyruvate:ferredoxin oxidoreductase (PFOR) catalyzes the oxidative decarboxylation of pyruvate to acetyl-CoA and CO(2). The catalytic proficiency of this enzyme for the reverse reaction, pyruvate synthase, is poorly understood. Conversion of acetyl-CoA to pyruvate links the Wood-Ljungdahl pathway of autotrophic CO(2) fixation to the reductive tricarboxylic acid cycle, which in these autotrophic anaerobes is the stage for biosynthesis of all cellular macromolecules. The results described here demonstrate that the Clostridium thermoaceticum PFOR is a highly efficient pyruvate synthase. The Michaelis-Menten parameters for pyruvate synthesis by PFOR are: V(max) = 1.6 unit/mg (k(cat) = 3.2 s(-1)), K(m)(Acetyl-CoA) = 9 micrometer, and K(m)(CO(2)) = 2 mm. The intracellular concentrations of acetyl-CoA, CoASH, and pyruvate have been measured. The predicted rate of pyruvate synthesis at physiological concentrations of substrates clearly is sufficient to support the role of PFOR as a pyruvate synthase in vivo. Measurements of its k(cat)/K(m) values demonstrate that ferredoxin is a highly efficient electron carrier in both the oxidative and reductive reactions. On the other hand, rubredoxin is a poor substitute in the oxidative direction and is inept in donating electrons for pyruvate synthesis.  相似文献   

17.
We present a robust second-generation BioScope: a system for continuous perturbation experiments. Firstly, the BioScope design parameters (i.e., pressure drop, overall oxygen (O2) and carbon dioxide (CO2) mass transfer, mean residence time distribution and plug flow characteristics) were evaluated. The average overall mass transfer coefficients were estimated to be 1.8E-5 m s(-1) for O2 and 0.34E-5 m s(-1) for CO2. It was determined that the O2/CO2 permeable membrane accounted for 75% and 95% of the overall resistance for O2 and CO2, respectively. The Peclet number (Pe) of the system was found to be >500 for liquid flow rates between 1 and 4 ml min(-1), ensuring plug flow characteristics. Secondly, steady-state intracellular metabolite concentrations obtained using direct rapid sampling from the fermentor were compared with those obtained by rapid sampling via the pre-perturbation sample port of the BioScope. With both methods the same metabolite levels were obtained. Thirdly, glucose perturbation experiments were carried out directly in the fermentor as well as in the BioScope, whereby steady-state Saccharomyces cerevisiae cells from a glucose/ethanol limited chemostat were perturbed by increasing the extracellular glucose concentration from 0.11 to 2.8 mM. Intracellular and extracellular metabolite levels were measured within a time window of 180 s. It was observed that the dynamic metabolite concentration profiles obtained from both perturbations were nearly the same, with the exception of the C4 metabolites of the TCA cycle, which might be due to differences in culture age.  相似文献   

18.
To optimize the aeration conditions for microalgal biomass production in a vertical flat-plate photobioreactor (VFPP), the effect of the aeration rate on biomass productivity was investigated under given conditions. Air enriched with 5% or 10% (v/v) CO(2) was supplied for the investigation at rates of 0.025-1 vvm. The CO(2) utilization efficiency, change of pH in the medium, and the optimum aeration rate were determined by evaluating biomass productivity. To investigate the VFPP mass transfer characteristics, the overall volumetric mass transfer coefficient, k(L)a, was evaluated for several different flat-plate sizes. Increasing the height of the VFPP could improve both the mass transfer of CO(2) and the illumination conditions, so this appeared to be a good method for scaling up. Based on a comparison of the k(L)a value at the optimum aeration rate with previously reported results, it was confirmed that the range of CO(2) concentration used in the experiments was cost-effective for mass culture.  相似文献   

19.
Macrophytic marine red algae are a diverse source of bioactive natural compounds. "Microplantlet" suspension cultures established from red algae are potential platforms for biosynthesis of these compounds, provided suitable bioreactor configurations for mass culture can be identified. The stirred tank bioreactor offers high rates of gas-liquid mass transfer, which may facilitate the delivery of the CO(2) in the aeration gas to the phototrophic microplantlet suspension culture. Therefore, the effects of impeller speed and CO(2) delivery on the long-term production of microplantlet biomass of the model red alga Agardhiella subulata was studied within a stirred tank photobioreactor equipped with a paddle blade impeller (D(i)/D(T) = 0.5). Nutrient medium replacement was required for sustained biomass production, and the biomass yield coefficient based on nitrate consumption was 1.08 +/- 0.09 g dry biomass per mmol N consumed. Biomass production went through two exponential phases of growth, followed by a CO(2) delivery limited growth phase. The CO(2)-limited growth phase was observed only if the specific growth rate in the second exponential phase of growth was at least 0.03 day(-)(1), the CO(2) delivery rate was less than 0.258 mmol CO(2) L(-)(1) culture h(-)(1), and the plantlet density was at least 10 g fresh mass L(-)(1). Increasing the aeration gas CO(2) partial pressure from 0.00035 to 0.0072 atm decreased the cultivation pH from 8.8 to 7.8, prolonged the second exponential phase of growth by increasing the CO(2) delivery rate, and also increased the photosynthetic oxygen evolution rate. Impeller speeds ranging from 60 to 250 rpm, which generated average shear rates of 2-10 s(-)(1), did not have a significant effect on biomass production rate. However, microplantlets cultivated in a stirred tank bioreactor ultimately assumed compact spherical shape, presumably to minimize exposure to hydrodynamic stress.  相似文献   

20.
Fermentations of Penicillium chrysogenum have been made with different CO(2) contents in the influent gas streams. The rheological behavior of the culture broth was found to be significantly changed by exposure to high levels of CO(2). This is attributed to the wide variation in the morphology of P. chrysogenum, from normal mycelia with long hyphae to roughly spherical pellets when subjected to high levels of CO(2). A correlation has been developed relating volumetric O(2) transfer coefficients, k(L)a, with the effective O(2) diffusion coefficients, D(e), and the apparent viscosities, mu(app), based on the results obtained in this study. The use of CO(2) as a potent means for altering the rheological properties of culture broths and consequently improving the O(2) transfer capabilities in penicillin fermentations was clearly demonstrated.  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