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 78 毫秒
1.
Summary The effect of pH on growth and lactic acid production ofLactobacillus helveticus was investigated in a continuous culture using supplemented whey ultrafiltrate. Maximum lactate productivity of 5 gl–1h–1 occurred at pH 5.5. Whey permeates concentrated up to four times were fermented using batch cultures. Maximum lactic acid concentration of 95 gl–1 was attained, but residual sugars indicated a possible limitation in growth factors.Nomenclature D Dilution rate [h–1] - X Biomass [gl–1] - Glu Glucose consentration [gl–1] - Gal Galactose consentration [gl–1] - S Substrate, Lactose consentration [gl–1] - P Product, Lactate consentration [gl–1] - Yp/s Yield, defined as P/S [gg–1] - ri Rate of synthesis or consumption of i [gl–1h–1]  相似文献   

2.
Previous studies have shown that the rate of formation of streptokinase, a secondary metabolite, in batch fermentation is proportional to the specific growth rate of the biomass, which in turn is inhibited by its substrate and the primary product (lactic acid). These kinetics suggest the suitability of fed-batch operation to increase the yield of streptokinase. A near-optimal feed policy has been calculated by the chemotaxis algorithm, and it shows a substrate feed rate decreasing nonlinearly and vanishing after 11 hours. This is followed by batch fermentation for a further 8 hours, at the end of which 12% more streptokinase is generated than by purely batch fermentation. Further improvements in productivity are possible.List of Symbols k dh–1 decay constant for active cells - k ph–1 decay constant for streptokinase - K Igl–1 inhibition constant for lactic acid - KS gl–1 inhibition constant for substrate - M gl–1 lactic acid concentration - P gl–1 streptokinase concentration - Q 1h–1 substrate feed rate - S gl–1 substrate concentration - S ingl–1 inlet concentration of substrate - t h time - t bh end-point of batch fermentation - t fh end-point of fed-batch fermentation - V l volume of broth in fermenter - V 0 l initial value of V (at t=0) - V ml maximum value of V - X gl–1 total biomass concentration - X agl–1 concentration of active biomass - Y MX yield coefficient for lactic acid from biomass - Y PX yield coefficient for streptokinase from biomass - Y XS yield coefficient for biomass from substrate Greek Letters h–1 specific growth rate of biomass - mh–1 maximum specific growth rate  相似文献   

3.
A fermentation medium based on millet (Pennisetum typhoides) flour hydrolysate and a four-phase feeding strategy for fed-batch production of baker's yeast,Saccharomyces cerevisiae, are presented. Millet flour was prepared by dry-milling and sieving of whole grain. A 25% (w/v) flour mash was liquefied with a thermostable 1,4--d-glucanohydrolase (EC 3.2.1.1) in the presence of 100 ppm Ca2+, at 80°C, pH 6.1–6.3, for 1 h. The liquefied mash was saccharified with 1,4--d-glucan glucohydrolase (EC 3.2.1.3) at 55°C, pH 5.5, for 2 h. An average of 75% of the flour was hydrolysed and about 82% of the hydrolysate was glucose. The feeding profile, which was based on a model with desired specific growth rate range of 0.18–0.23 h–1, biomass yield coefficient of 0.5 g g–1 and feed substrate concentration of 200 g L–1, was implemented manually using the millet flour hydrolysate in test experiments and glucose feed in control experiments. The fermentation off-gas was analyzed on-line by mass spectrometry for the calculation of carbon dioxide production rate, oxygen up-take rate and the respiratory quotient. Off-line determination of biomass, ethanol and glucose were done, respectively, by dry weight, gas chromatography and spectrophotometry. Cell mass concentrations of 49.9–51.9 g L–1 were achieved in all experiments within 27 h of which the last 15 h were in the fedbatch mode. The average biomass yields for the millet flour and glucose media were 0.48 and 0.49 g g–1, respectively. No significant differences were observed between the dough-leavening activities of the products of the test and the control media and a commercial preparation of instant active dry yeast. Millet flour hydrolysate was established to be a satisfactory low cost replacement for glucose in the production of baking quality yeast.Nomenclature C ox Dissolved oxygen concentration (mg L–1) - CPR Carbon dioxide production rate (mmol h–1) - C s0 Glucose concentration in the feed (g L–1) - C s Substrate concentration in the fermenter (g L–1) - C s.crit Critical substrate concentration (g L–1) - E Ethanol concentration (g L–1) - F s Substrate flow rate (g h–1) - i Sample number (–) - K e Constant in Equation 6 (g L–1) - K o Constant in Equation 7 (mg L–1) - K s Constant in Equation 5 (g L–1) - m Specific maintenance term (h–1) - OUR Oxygen up-take rate (mmol h–1) - q ox Specific oxygen up-take rate (h–1) - q ox.max Maximum specific oxygen up-take rate (h–1) - q p Specific product formation rate (h–1) - q s Specific substrate up-take rate (g g–1 h–1) - q s.max Maximum specific substrate up-take rate (g g–1 h–1) - RQ Respiratory quotient (–) - S Total substrate in the fermenter at timet (g) - S 0 Substrate mass fraction in the feed (g g–1) - t Fermentation time (h) - V Instantaneous volume of the broth in the fermenter (L) - V 0 Starting volume in the fermenter (L) - V si Volume of samplei (L) - x Biomass concentration in the fermenter (g L–1) - X 0 Total amount of initial biomass (g) - X t Total amount of biomass at timet (g) - Y p/s Product yield coefficient on substrate (–) - Y x/e Biomass yield coefficient on ethanol (–) - Y x/s Biomass yield coefficient on substrate (–) Greek letters Moles of carbon per mole of yeast (–) - Moles of hydrogen atom per mole of yeast (–) - Moles of oxygen atom per mole of yeast (–) - Moles of nitrogen atom per mole of yeast (–) - Specific growth rate (h–1) - crit Critical specific growth rate (h–1) - E Specific ethanol up-take rate (h–1) - max.E Maximum specific ethanol up-take rate (h–1)  相似文献   

4.
Summary Unidirectional 22Na+ and 36Cl fluxes were determined in short-circuited, stripped rumen mucosa from sheep by using the Ussing chamber technique. In both CO2/HCO 3 -containing and CO2/HCO 3 -free solutions, replacement of gluconate by short-chain fatty acids (SCFA, 39 mM) significantly enhanced mucosal-toserosal Na+ absorption without affecting the Cl transport in the same direction. Short-chain fatty acid stimulation of Na+ transport was at least partly independent of Cl and could almost completely be abolished by 1 mM mucosal amiloride, while stimulation of Na+ transport was enhanced by lowering the mucosal pH from 7.3 to 6.5. Similar to the SCFA action, raising the PCO2 in the mucosal bathing solution led to an increase in the amiloride-sensitive mucosal-to-serosal Na+ flux. Along with its effect on sodium transport, raising the PCO2 also stimulated chloride transport. The results are best explained by a model in which undissociated SCFA and/or CO2 permeate the cell membrane and produce a raise in intracellular H+ concentration. This stimulates an apical Na+/H+ exchange, leading to increased Na+ transport. The stimulatory effect of CO2 on Cl transport is probably mediated by a Cl/HCO 3 exchange mechanism in the apical membrane. Binding of SCFA anions to that exchange as described for the rat distal colon (Binder and Mehta 1989) probably does not play a major role in the rumen.Abbreviations DIDS 4,4'-diisothiocyanatostilbene-2,2'-disulfonic acid - G t transepithelial conductance (mS·cm-2) - HSCFA undissociated short-chain fatty acids - J ms mucosal-to-serosal flux (Eq · cm-2 · h-1) - J net net flux (Eq · cm-2 · h-1) - J sm serosal-to-mucosal flux (Eq · cm-2 · h-1) - PD transepithelial potential difference (mV) - SCFA dissociated short-chain fatty acids - SCFA short-chain fatty acids  相似文献   

5.
Summary The influence of temperature on the growth of the theromophilic Bacillus caldotenax was investigated using chemostat techniques and a chemically defined minimal medium. All determined growth constants, that is maximal specific growth rate, yield and maintenance, were temperature dependent. It was striking that the very large maintenance requirement was about 10 times higher than for mesophilic cells under equivalent conditions. A death rate, which was very substantial at optimal and supraoptimal growth temperatures, was estimated by comparing the maintenance for substrate and oxygen. There was no indication for a thermoadaptation as postulated by Haberstich and Zuber (1974).Symbols D Dilution rate (h–1) - Dc=max Critical dilution rate (h–1) - E Temperature characteristic (J mol–1) - k Organism constant - kd Death rate coefficient (h–1) - km Maintenance substrate coefficient estimated from MO (h–1) - MO Maintenance respiration, mmol O2 per g dry biomass and h (mmol g–1h–1) - MO Maintenance respiration, taking kd into account - mS Maintenance substrate coefficient, g glucose per g dry biomass and h (h–1) - OD Optical density at 546 nm - QO2 Specific O2-uptake rate (mmol g–1h–1) - Q O2 V Specific O2-uptake rate for viable portion of biomass (mmol g–1 h–1) - QS Specific glucose uptake rate (h–1) - Q S V Specific glucose uptake rate for viable portion of biomass (h–1) - R Gas constant 8.28 J mol–1K–1 - S Substrate concentration in reactor (g l–1) - SO Influent substrate concentration (g l–1) - Tmax Maximal growth temperature (°C) - Tmin Minimal growth temperature (°C) - X Dry biomass (g l–1) - XtOt=X Dry biomass containing dead and viable cells - Xv Viable portion of biomass - Y O m Potential yield for O2 corrected for maintenance respiration (g mol–1) - Y S m Potential yield for substrate corrected for maintenance requirement, g biomass per g glucose (–) - Specific growth rate (h–1) - max Maximal specific growth rate (h–1)  相似文献   

6.
Summary Conditions for the production of microbial uricase byCandida utilis were studied. For the selected strain, hypoxanthine proved to be the most effective inducer of uricase formation. The highest values of biomass as well as uricase activity in the mechanically agitated fermentor were obtained under the following conditions: 50 h, rotation impeller speed 7 s–1, air flow rate 1.25×10–5 m3s–1, concentration of inducer 0.1%.List of symbols b width of baffle, m - c length of baffle, m - D diameter of cylindrical fermentor, m - d diameter of impeller, m - d 1 diameter of impeller disc, m - Fr m impeller Froud number - g gravitional acceleration, ms–2 - H height of batch surface above bottom, m - H 2 height of impeller disc above bottom, m - h height of impeller blade, m - Kp g flow rate number - L length of impeller blade, m - N rotational speed of impeller, s–1 - Re m impeller Reynolds number - T time, h - V volume of batch, m3 - V g air (gas) flow rate, m3s–1 - x mass fraction of the dry matter of cells - x 0 initial value of the mass fraction of the dry matter of cells - r volume fraction of the dry matter of cells - <eta<1 viscosity of pure liquid, Pa s - viscosity of batch (suspension of microbial suspension), Pa s - density of batch, kg m–3  相似文献   

7.
Diel vertical migration and feeding cycles of adult female Metridia gerlachei in the upper 290 m of a 335-m water column were measured during a total of 65 h in two periods of early summer (Dec 20–21 and Dec 25–26, 1991). Samples collected in eight depth strata by 35 MOCNESS tows (333-m mesh) were analyzed for abundance and mean individual gut pigment content. Most of the copepod population was concentrated in a 50-m depth interval at all times. Feeding began simultaneously with nocturnal ascent from a depth of 200–250 m at 18:00 h (local time), when the relative change in ambient light intensity was greatest. Ingestion rate increased exponentially (ki = 0.988 h–1) at double the gut evacuation rate (ke = 0.488 h–1) as the population moved upward at 22.3–26.5 m h–1 through increasing concentrations of particulate chlorophyll-a. Although the bulk of the population did not move to depths shallower than 50 m, and began its downward migration at a rate of 20.8–31.7 mh–1 in complete darkness, individual females continued to make brief excursions into chlorophyll-rich surface waters (4–8 g l–1) during the first few hours of population descent. Ingestion rate diminished abruptly by one order of magnitude (ki = 0.068 h–1) at dawn ( 0330 h). Within four more hours, the population had reached its daytime depth and gut pigment content remained constant at a minimum value until the next migration cycle. No feeding appeared to take place at depth during the day. Ingestion by M. gerlachei females removed < 4% of daily primary production, with only 20% of this amount being removed from surface waters by active vertical transport.  相似文献   

8.
Studies in tower reactors with viscous liquids on flow regime, effective shear rate, liquid mixing, gas holdup and gas/ liquid mass transfer (k La) are reviewed. Additional new data are reported for solutions of glycerol, CMC, PAA, and xanthan in bubble columns with diameters of 0.06, 0.14 and 0.30 m diameter. The wide variation of the flow behaviour index (1 to 0.18) allows to evaluate the effective shear rate due to the gas flow. New dimensionless correlations are developed based on the own and literature data, applied to predict k La in fermentation broths, and compared to other reactor types.List of Symbols a(a) m–1 specific interfacial area referred to reactor (liquid) volume - Bo Bond number (g D c 2 L/) - c L(c L * ) kmol m–3 (equilibrium) liquid phase oxygen concentration - C coefficient characterising the velocity profile in liquid slugs - C s m–1 coefficient in Eq. (2) - d B(dvs) m bubble diameter (Sauter mean of d B) - d 0 m diameter of the openings in the gas distributor plate - D c m column diameter - D L m2s–1 diffusivity - E L(EW) m2 s–1 dispersion coefficient (in water) - E 2 square relative error - Fr Froude number (u G/(g Dc)0.5) - g m s–2 gravity acceleration - Ga Gallilei number (g D c 3 L 2 / eff 2 ) - h m height above the gas distributor the gas holdup is characteristic for - k Pasn fluid consistency index (Eq. 1) - k L m s–1 liquid side mass transfer coefficient - k La(kLa) s–1 volumetric mass transfer coefficient referred to reactor (liquid) volume - L m dispersion height - n flow behaviour index (Eq. 1) - P W power input - Re liquid slug Reynolds number ( L(u G +u L) D c/eff) - Sc Schmidt number ( eff/( L D L )) - Sh Sherwood number (k La D c 2 /DL) - t s time - u B(usw) m s–1 bubble (swarm) rise velocity - u G(uL) m s–1 superficial gas (liquid) velocity - V(VL) m3 reactor (liquid) volume Greec Symbols W m–2 K–1 heat transfer coefficient - y(y eff) s–1 (effective) shear rate - G relative gas holdup - s relaxation time of viscoelastic liquid - L(eff) Pa s (effective) liquid viscosity (Eq. 1) - L kg m–3 liquid density - N/m surface tension  相似文献   

9.
Summary Deficiency of inorganic phosphate caused the hyper production of invertase and the derepression of acid phosphatase in a continuous culture ofSaccharomyces carlsbergensis. The specific invertase activity was 40,000 enzyme units per g dry cell weight at a dilution rate lower than 0.05 h–1 with a synthetic glucose medium of which the molecular ratio of KH2PO4 to glucose was less than 0.006. This activity is eight fold higher than in a batch growth and 1.5 fold as much as the highest enzyme activity observed so far in a glucose-limited continuous culture.For the hyper production of invertase, it is necessary to culture the yeast continuously by keeping the Nyholm's conservative inorganic phosphate concentration at less than 0.2 m mole per g dry weight cell. The derepression of acid phosphatase brought about by phosphate deficiency, was similar in both batch and continuous cultures.Nomenclature D dilution rate of continuous culture (h–1) - Ei invertase concentration in culture (enzyme unit l–1) - Ep acid phosphatase concentration in culture (enzyme unit l–1) - P inorganic phosphate concentration in culture (mM) - S glucose concentration in culture (mM) - X cell concentration in culture (g dry weight cell l–1) Greek Letter specific rate of growth (h–1) Suffix f feed - 0 initial value  相似文献   

10.
A family of 10 competing, unstructured models has been developed to model cell growth, substrate consumption, and product formation of the pyruvate producing strain Escherichia coli YYC202 ldhA::Kan strain used in fed-batch processes. The strain is completely blocked in its ability to convert pyruvate into acetyl-CoA or acetate (using glucose as the carbon source) resulting in an acetate auxotrophy during growth in glucose minimal medium. Parameter estimation was carried out using data from fed-batch fermentation performed at constant glucose feed rates of qVG=10 mL h–1. Acetate was fed according to the previously developed feeding strategy. While the model identification was realized by least-square fit, the model discrimination was based on the model selection criterion (MSC). The validation of model parameters was performed applying data from two different fed-batch experiments with glucose feed rate qVG=20 and 30 mL h–1, respectively. Consequently, the most suitable model was identified that reflected the pyruvate and biomass curves adequately by considering a pyruvate inhibited growth (Jerusalimsky approach) and pyruvate inhibited product formation (described by modified Luedeking–Piret/Levenspiel term).List of symbols cA acetate concentration (g L–1) - cA,0 acetate concentration in the feed (g L–1) - cG glucose concentration (g L–1) - cG,0 glucose concentration in the feed (g L–1) - cP pyruvate concentration (g L–1) - cP,max critical pyruvate concentration above which reaction cannot proceed (g L–1) - cX biomass concentration (g L–1) - KI inhibition constant for pyruvate production (g L–1) - KIA inhibition constant for biomass growth on acetate (g L–1) - KP saturation constant for pyruvate production (g L–1) - KP inhibition constant of Jerusalimsky (g L–1) - KSA Monod growth constant for acetate (g L–1) - KSG Monod growth constant for glucose (g L–1) - mA maintenance coefficient for growth on acetate (g g–1 h–1) - mG maintenance coefficient for growth on glucose (g g–1 h–1) - n constant of extended Monod kinetics (Levenspiel) (–) - qV volumetric flow rate (L h–1) - qVA volumetric flow rate of acetate (L h–1) - qVG volumetric flow rate of glucose (L h–1) - rA specific rate of acetate consumption (g g–1 h–1) - rG specific rate of glucose consumption (g g–1 h–1) - rP specific rate of pyruvate production (g g–1 h–1) - rP,max maximum specific rate of pyruvate production (g g–1 h–1) - t time (h) - V reaction (broth) volume (L) - YP/G yield coefficient pyruvate from glucose (g g–1) - YX/A yield coefficient biomass from acetate (g g–1) - YX/A,max maximum yield coefficient biomass from acetate (g g–1) - YX/G yield coefficient biomass from glucose (g g–1) - YX/G,max maximum yield coefficient biomass from glucose (g g–1) - growth associated product formation coefficient (g g–1) - non-growth associated product formation coefficient (g g–1 h–1) - specific growth rate (h–1) - max maximum specific growth rate (h–1)  相似文献   

11.
Summary The nonsporulating extreme thermophile Thermus thermophilus was grown in continuous culture at dilution rates up to 2.65 h–1 at 75°C and pH 6.9 on complex medium. Concomitantly very low yield (Y=0.12 g cell dry weight g–1 utilized organic carbon) and incomplete substrate utilization (always less than 45%) were found. In batch cultures T. thermophilus could be grown with max =h–1, in shake flasks only with max =h–1 with the same low yield and incomplete substrate utilization. Stable steady states at 84C and 45°C were realized at a dilution rate of 0.3 h–1 whereas at 86°C and 40°C no growth could be detected. Artefacts arising from wall growth (in bioreactors) or improper materials must be ruled out. Inhibition of growth by organic substrates was demonstrated at low concentrations: a decrease in the yield obtained was found when more than 0.7 gl–1 of meat extract were supplied in the medium. The maintenance requirement for oxygen is potentially very high and was determined to be 10 to 15 mmol g–1 h–1.  相似文献   

12.
Summary Submerged batch cultivation under controlled environmental conditions of pH 3.8, temperature 30°C, and KLa200 h–1 (above 180 mMO2 l –1 h–1 oxygen supply rate) produced a maximum (12.0 g·l –1) SCP (Candida utilis) yield on the deseeded nopal fruit juice medium containing C/N ratio of 7.0 (initial sugar concentration 25 g·l –1) with a yield coefficient of 0.52 g cells/g sugar. In continuous cultivation, 19.9 g·l –1 cell mass could be obtained at a dilution rate (D) of 0.36 h–1 under identical environmental conditions, showing a productivity of 7.2 g·l –1·h–1. This corresponded to a gain of 9.0 in productivity in continuous culture over batch culture. Starting with steady state values of state variables, cell mass (CX–19.9 g·l –1), limiting nutrient concentration (Cln–2.5 g·l –1) and sugar concentration (CS–1.5 g·l –1) at control variable conditions of pH 3.8, 30°C, and KLa 200 h–1 keeping D=0.36 h–1 as reference, transient response studies by step changes of these control variables also showed that this pH, temperature and KLa conditions are most suitable for SCP cultivation on nopal fruit juice. Kinetic equations obtained from experimental data were analysed and kinetic parameters determined graphically. Results of SCP production from nopal fruit juice are described.Nomenclature Cln concentration of ammonium sulfate (g·l –1) - CS concentration of total sugar (g·l –1) - CX cell concentration (g·l –1) - D dilution rate (h–1) - Kln Monod's constant (g·l –1) - m maintenance coefficient (g ammonium sulfate cell–1 h–1) - m(S) maintenance coefficient (g sugar g cell–1 h–1) - t time, h - Y yield coefficient (g cells/g ammonium sulfate) - Ym maximum of Y - YS yield coefficient based on sugar consumed (g cells · g sugar–1) - YS(m) maximum value of YS - µm maximum specific growth rate constant (h–1)  相似文献   

13.
Animal coat color and radiative heat gain: A re-evaluation   总被引:1,自引:0,他引:1  
Summary Thermal resistance and heat gain from simulated solar radiation were measured over a range of wind velocities in black and white pigeon plumages. Plumage thermal resistance averaged 39% (feathers depressed) or 16% (feathers erected) of that of an equivalent depth of still air. Feather erection increased plumage depth four-fold and increased plumage thermal resistance about 56%. At low wind speeds, black plumages acquired much greater radiative heat loads than did white plumages. However, associated with the greater penetration of radiation into light than dark plumages, the radiative heating of white plumages is affected less by convective cooling than is that of black plumages. Thus, the heat loads of black and white plumages converge as wind speed is increased. This effect is most prominent in erected plumages, where at wind speeds greater than 3 ms–1 black plumages acquire lower radiative heat loads than do white plumages. These results suggest that animals with dark-colored coats may acquire lower heat loads under ecologically realistic conditions than those forms with light-colored coats. Thus, the dark coat colors of a number of desert species and the white coat color of polar forms may be thermally advantageous.These results are used to test a new general model that accounts for effects of radiation penetration into a fur or feather coat upon an animal's heat budget. Even using simplifying assumptions, this model's predictions closely match measured values for plumages with feathers depressed (the typical state). Predictions using simplifying assumptions are less accurate for erected plumages. However, the model closely predicts empirical data for erected white plumages if one assumption is obviated by additional measurements. Data are not sufficient to judge whether this is also the case for erected black plumages.List of Symbols A body surface area (m2) - a L long-wave absorptivity of coat - a s short-wave absorptivity of coat - d characteristic dimension (m) - E evaporative water loss (kg m–2 s–1) - h coat thermal conductance (W m–2 °C–1) - k convection constant (s1/2 m–1) - l coat thickness (m) - L i long-wave irradiance at coat surface (W m–2) - M metabolic heat production (W m–2) - m body mass (kg) - P plumage mass (kg) - p probability per unit coat depth that a penetrating ray will strike a coat element (m–1) - q(Z) radiation absorbed at level z (W m–2) - R abs radiation absorbed by animal (W m–2) - r e external resistance to convective and radiative heat transfer (s m–1) - r Ha boundary layer resistance to convective heat transfer (s m–1) - r Hb whole-body thermal resistance (s m–1) - r Hc coat (plumage) thermal resistance (s m–1) - r Ht tissue thermal resistance (s m–1) - r s apparent resistance to radiative heat transfer (s m–1) - r(Z) thermal resistance from level z to coat surface (s m–1) - S i short-wave irradiance at coat surface (W m–2) - S radiant flux going toward skin surface (W m–2) - S + radiant flux going away from skin surface (W m–2) - T a air temperature (°C) - T b core body temperature (°C) - T e equivalent black-body temperature (°C) - T e air temperature plus temperature increment due to longwave radiation (°C) - u wind velocity (m s–1) - V heat load on animal from short-wave radiation (W m–2) - z depth within coat (m) - short-wave absorptivity of individual hairs or feather elements - emissivity - {ie211-1} - latent heat of vaporization of water (J kg–1) - short-wave reflectivity of individual hairs or feather elements - {ie211-2} short-wave reflectivity of coat - {ie212-1} short-wave reflectivity of skin - c p volumetric specific heat of air (J m–3 °C–1) - Stefan-Boltzmann constant (W m–2 °K–4) - short-wave transmissivity of individual hairs or feather elements - {ie212-2} short-wave transmissivity of coat  相似文献   

14.
Summary The kinetics ofBordetella pertussis growth was studied in a glutamate-limited continuous culture. Growth kinetics corresponded to Monod's model. The saturation constant and maximum specific growth rate were estimated as well as the energetic parameters, theoretical yield of cells and maintenance coefficient. Release of pertussis toxin (PT) and lipopolysaccharide (LPS) were growth-associated. In addition, they showed a linear relationship between them. Growth rate affected neither outer membrane proteins nor the cell-bound LPS pattern.Nomenclature X cell concentration (g L–1) - specific growth rate (h–1) - m maximum specific growth rate (h–1) - D dilution rate (h–1) - S concentration of growth rate-limiting nutrient (glutamate) (mmol L–1 or g L–1) - Ks substrate saturation constant (mol L–1) - ms maintenance coefficient (g g–1 h–1) - Yx/s theoretical yield of cells from glutamate (g g–1) - Yx/s yield of cells from glutamate (g g–1) - YPT/s yield of soluble PT from glutamate (mg g–1) - YKDO/s yield of cell-free KDO from glutamate (g g–1) - YPT/x specific yield of soluble PT (mg g–1) - YKDO/x specific yield of cell-free KDO (g g–1) - qPT specific soluble PT production rate (mg g–1 h–1) - qKDO specific cell-free KDO production rate (g g–1 h–1)  相似文献   

15.
The kinetics of the release of chitinolytic activity (endochitinase EC 3.2.1.14, \-N-acetyglucosaminidase EC 3.2.1.30) by a yeast cell wall lytic Arthrobacter species was studied. The organism was cultivated on yeast cell wall, mycelium of Trichoderma reesei, colloidal chitin, N-acetylglucosamine, glucosamine and mixtures with acetate. With the exception of yeast cell wall, these substrates were used as the sole source of carbon and nitrogen. The growth on colloidal chitin (0.5%) proceeded at a maximum specific growth rate (umax) of 0.23 h–1 and yielded 2700 mU1–1 chitinase. Yeast cell wall and mycelium of T. reesei supported more rapid growth (max = 0.30 h–1 and 0.25 h–1 respectively) but yielded reduced chitinase activity (565 mUl–1 and 700 mUl–1). The growth rate on glucosamine (max = 0.24 h–1) was reduced when this was mixed with acetate (max = 0.12 h–1), whereas the enzyme yield was increased from 720 mUl–1 to 960 mUl–1. The same effect on growth rate was observed with glucose and equimolar mixtures of glucose and acetate, indicating a strong impact of the organic acid on carbohydrate transport or metabolism. The growth of adapted cells on N-acetylglucosamine was comparable to that observed on an equimolar mixture of glucosamine and acetate, indicating that N-acetylglucosamine is rapidly hydrolysed by adapted cells.  相似文献   

16.
Summary The influence of different operational parameters, such as the dilution rate (D) and the bleeding rate (B), in the production of a flocculent strain ofLactobacillus plantarum was studied. The effect of the dilution rate was demonstrated to be related to the lactic acid concentration inside the reactor. The effect of the bleeding rate was shown to be critical in the stabilization of the operation (due to a better pH control). It also allowed a continuous recovery of cells outside the reactor. Viability testing of the lactic starter cultures showed that operation with cell purge increased the viability of the starter cultures obtained.Nomenclature B Bleeding rate, h–1 - D Dilution rate, h–1 - F Feed flow rate, L h–1 - I Feed velocity, m h–1 - Specific growth rate, h–1 - v Lactic acid specific productivity, g g–1 h–1 - P Product concentration (lactic acid), g L–1 - P out Product concentration leaving the system, g L–1 - Q b Bleeding flow rate, L h–1 - R Recirculation velocity, m h–1 - S Substract concentration, g L–1 - t Time, h - T p Time of ascensional flow (length of the column/total ascensional velocity), h - T r Residence time (1/D), h - V Volume of the reactor, L - X Cell concentration, g L–1 - X out Cell concentration leaving the system, g L–1  相似文献   

17.
Rhodospirillum rubrum was grown continuously and photoheterotrophically under light limitation using a cylindrical photobioreactor in which the steady state biomass concentration was varied between 0.4 to 4 kg m–3 at a constant radiant incident flux of 100 W m–2. Kinetic and stoichiometric models for the growth are proposed. The biomass productivities, acetate consumption rate and the CO2 production rate can be quantitatively predicted to a high level of accuracy by the proposed model calculations. Nomenclature: C X, biomass concentration (kg m–3) D, dilution rate (h–1) Ea, mean mass absorption coefficient (m2 kg–1) I , total available radiant light energy (W m–2) K, half saturation constant for light (W m–2) R W, boundary radius defining the working illuminated volume (m) r X, local biomass volumetric rate (kg m–3 h–1) <r X>, mean volumetric growth rate (kg m–3 h–1) V W, illuminated working volume in the PBR (m–3). Greek letters: , working illuminated fraction (–) M, maximum quantum yield (–) bar, mean energetic yield (kg J–1).  相似文献   

18.
Summary Some environmental affects on cell aggregation described in the literature are briefly summarized. By means of a biomass recirculation culture (Contact system), using the yeast Torulopsis glabrata, the aggregation behavior of cells in static and in dynamic test systems is described. Sedimentation times required to obtain 50 g · l–1 yeast dry matter in static systems were always higher than in dynamic ones.In addition to, influencing the biomass yield, the specific growth rate of the yeast also affected cell aggregation. The specific growth rate and therefore the aggregation could be regulated by the biomass recirculation rate as well as by the sedimenter volume.Abbreviations fo Overflow flow rate (l·h–1) - fR Recycle flow rate (l·h–1) - ft0t Total flow rate through the fermenter (l·h–1) - g Gram - h Hour - DR Fermenter dilution rate due to recycle (h–1) - DS Fermeter dilution rate due to substrate (h–1) - Dtot Total fermenter dilution rate (h–1) - l Liter - Specific growth rate (h–1) - PF Fermenter productivity (g·l–1·h–1) - PFS Overall productivity (g·l–1·h–1) - RpM Rates per minute - RS Residual sugar content in the effluent with respect to the substrate concentration (%) - Y Yield of biomass with respect to sugar concentration (%) - Sed 50 Sedimentation time to reach a YDM of 50 g·l–1 (min) - V Volume (l) - VF Fermenter volume (l) - VSed Sedimenter volume (l) - VVM Volumes per volume and minute - XF YDM in the fermenter (g·l–1) - XF YDM in the recycle (g·l–1) - XS Yeast dry matter due to substrate concentration (g·l–1) - YDM Yeast dry matter (g·l–1)  相似文献   

19.
Summary A strain of the yeast Lipomyces kononenkoae which converted starch into SCP with a high yield, produced three extracellular amylases which were purified from the culture fluid by Ficoll concentration, dialysis, isopropanol precipitation and DE-cellulose chromatography: an -amylase, a glucoamylase and a debranching transferase. The latter transferred -1,6-glucosyl units from panose to glucose forming maltose and appeared to have some debranching activity on amylopectin. The -amylase had the following properties: MW 38000 daltons; no effect of added calcium ions on activity; optimum temperature and pH for activity around 40°C and pH 5.5; H and S of heat inactivation 24360 cal mol–1 and 29.2 cal deg–1 mol–1; range of pH stability pH 4–6.5; pI=7.1; final low molecular weight products of starch hydrolysis, maltose and glucose; Km (40°C, pH 5.5) for starch 2.7 gl–1, for maltotriose 109 gl–1; uncompetitive inhibition by maltose with Ki (40°C, pH 5.5) 29.5 gl–1. The glucoamylase had the following properties: MW 81500 daltons; optimum temperature and pH for activity around 50°C and pH 4.5: H and S of heat inactivation 20400 cal mol–1 and 17.7 cal deg–1 mol–1; range of pH stability pH 4–6.5; pI=6.1; Km (30°C, pH 4.5) for soluble starch 16.2 gl–1, for maltose 0.36 gl–1, for p-nitrophenyl--D-glucoside 0.35 gl–1; competitive inhibition by glucose with Ki (30°C, pH 4.5) 4.7 gl–1.  相似文献   

20.
We studied the effects of atmospheric CO2 enrichment (280, 420 and 560 l CO2 l–1) and increased N deposition (0,30 and 90 kg ha–1 year–1) on the spruce-forest understory species Oxalis acetosella, Homogyne alpina and Rubus hirtus. Clones of these species formed the ground cover in nine 0.7 m2 model ecosystems with 5-year-old Picea abies trees (leaf area index of approx 2.2). Communities grew on natural forest soil in a simulated montane climate. Independently of N deposition, the rate of light-saturated net photosynthesis of leaves grown and measured at 420 l CO2 l–1 was higher in Oxalis and in Homogyne, but was not significantly different in Rubus compared to leaves grown and measured at the pre-industrial CO2 concentration of 280 l l–1. Remarkably, further CO2 enrichment to 560 l l–1 caused no additional increase of CO2 uptake. With increasing CO2 supply concentrations of non-structural carbohydrates in leaves increased and N concentrations decreased in all species, whereas N deposition had no significant effect on these traits. Above-ground biomass and leaf area production were not significantly affected by elevated CO2 in the more vigorously growing species O. acetosella and R. hirtus, but the slow growing H. alpina produced almost twice as much biomass and 50% more leaf area per plant under 420 l CO2 l–1 compared to 280 l l–1 (again no further stimulation at 560 l l–1). In contrast, increased N addition stimulated growth in Oxalis and Rubus but had no effect on Homogyne. In Oxalis (only) biomass per plant was positively correlated with microhabitat quantum flux density at low CO2, but not at high CO2 indicating carbon saturation. On the other hand, the less shade-tolerant Homogyne profited from CO2 enrichment at all understory light levels facilitating its spread into more shady micro-habitats under elevated CO2. These species-specific responses to CO2 and N deposition will affect community structure. The non-linear responses to elevated CO2 of several of the traits studied here suggest that the largest responses to rising atmospheric CO2 are under way now or have already occurred and possible future responses to further increases in CO2 concentration are likely to be much smaller in these understory species.  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