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 15 毫秒
1.
Benzene, toluene, ethylbenzene and xylene (BTEX) substrate interactions for a mesophilic (25°C) and thermophilic (50°C) toluene-acclimatized composted pine bark biofilter were investigated. Toluene, benzene, ethylbenzene, o-xylene, m-xylene and p-xylene removal efficiencies, both individually and in paired mixtures with toluene (1:1 ratio), were determined at a total loading rate of 18.1 g m–3 h–1 and retention time ranges of 0.5–3.0 min and 0.6–3.8 min for mesophilic and thermophilic biofilters, respectively. Overall, toluene degradation rates under mesophilic conditions were superior to degradation rates of individual BEX compounds. With the exception of p-xylene, higher removal efficiencies were achieved for individual BEX compounds compared to toluene under thermophilic conditions. Overall BEX compound degradation under mesophilic conditions was ranked as ethylbenzene >benzene >o-xylene >m-xylene >p-xylene. Under thermophilic conditions overall BEX compound degradation was ranked as benzene >o-xylene >ethylbenzene >m-xylene >p-xylene. With the exception of o-xylene, the presence of toluene in paired mixtures with BEX compounds resulted in enhanced removal efficiencies of BEX compounds, under both mesophilic and thermophilic conditions. A substrate interaction index was calculated to compare removal efficiencies at a retention time of 0.8 min (50 s). A reduction in toluene removal efficiencies (negative interaction) in the presence of individual BEX compounds was observed under mesophilic conditions, while enhanced toluene removal efficiency was achieved in the presence of other BEX compounds, with the exception of p-xylene under thermophilic conditions.  相似文献   

2.
Removal of toluene in waste gases using a biological trickling filter   总被引:12,自引:0,他引:12  
The removal of toluene from waste gas was studied in a trickling biofilter. A high level of water recirculation (4.7 m h–1) was maintained in order to keep the liquid phase concentration constant and to achieve a high degree of wetting. For loads in the range from 6 to 150 g m–3 h–1 the maximum volumetric removal rate (elimination capacity) was 35±10 g m–3 h–1, corresponding to a zero order removal rate of 0.11±0.03 g m–2 h–1 per unit of nominal surface area. The surface removal was zero order above the liquid phase concentrations of approximately 1.0 g m–3, corresponding to inlet gas concentrations above 0.7–0.8 g m–3. Below this concentration the surface removal was roughly of first order. The magnitude of the first order surface removal rate constant, k1A , was estimated to be 0.08–0.27 m h–1 (k1A a=24–86 h–1). Near-equilibrium conditions existed in the gas effluent, so mass transfer from gas to liquid was obviously relatively fast compared to the biological degradation. An analytical model based on a constant liquid phase concentration through the trickling filter column predicts the effluent gas concentration and the liquid phase concentration for a first and a zero order surface removal. The experimental results were in reasonable agreement with a very simple model valid for conditions with an overall removal governed by the biological degradation and independent of the gas/liquid mass transfer. The overall liquid mass transfer coefficient, KLa, was found to be a factor 6 higher in the system with biofilm compared to the system without. The difference may be explained by: 1. Difference in the wetting of the packing material, 2. Mass transfer occurring directly from the gas phase to the biofilm, and 3. Enlarged contact area between the gas phase and the biofilm due to a rough biofilm surface.  相似文献   

3.
Summary The growth parameters ofPenicillium cyclopium have been evaluated in a continuous culture system for the production of fungal protein from whey. Dilution rates varied from 0.05 to 0.20 h–1 under constant conditions of temperature (28°C) and pH (3.5). The saturation coefficients in the Monod equation were 0.74 g l–1 for lactose and 0.14 mg l–1 for oxygen, respectively. For a wide range of dilution rates, the yield was 0.68 g g–1 biomass per lactose and the maintenance coefficient 0.005 g g–1 h–1 lactose per biomass, respectively. The maximum biomass productivity achieved was 2 g l–1 h–1 biomass at dilution rates of 0.16–0.17 h–1 with a lactose concentration of 20 g l–1 in the feed. The crude protein and total nucleic acid contents increased with a dilution rate, crude protein content varied from 43% to 54% and total nucleic acids from 6 to 9% in the range of dilution rates from 0.05 to 0.2 h–1, while the Lowry protein content was almost constant at approximately 37.5% of dry matter.Nomenclature (mg l–1) Co initial concentration of dissolved oxygen - (h–1) D dilution rate - (mg l–1) K02 saturation coefficient for oxygen - (g l–1) Ks saturation coefficient for substrate - (g g–1 h–1) lactose per biomass) m maintenance energy coefficient - (mM g–1 h–1O2 per biomass) Q02 specific oxygen uptake rate - (g l–1) S residual substrate concentration at steady state - (g l–1) So initial substrate concentration in feed - (min) t1/2 time when Co is equal to Co/2 - (g l–1) X biomass concentration - (g l–1) X biomass concentration at steady state - (g g–1 biomass per lactose) YG yield coefficient for cell growth - (g g–1 biomass per lactose) Yx/s overall yield coefficient - (h–1) specific growth rate  相似文献   

4.
ABSTRACT

A laboratory-scale biofilter unit packed with a mixture of compost, sugarcane bagasse, and granulated activated carbon (GAC) in the ratio of 55:30:15 by weight was used for a biofiltration study of air stream containing benzene, toluene, ethylbenzene, and o-xylene (BTEX). The effect of superficial velocity on mass transfer coefficient for the packing was studied by maintaining gas flow rates of 3, 4, 5, 6, and 8 L min?1 for inlet concentrations of 0.1, 0.4, and 0.8 g m?3 for each of benzene, toluene, ethylbenzene, and o-xylene. The maximum elimination capacity was found to be 20.92, 22.72, 20.73, and 18.94 g m?3 h?1 for BTEX, respectively, for stated flow rates. Removal efficiency of BTEX decreased from 99% to 71% for increasing inlet concentration from 0.1 to 0.8 g m?3. Gas film mass transfer coefficient predicted by modified Onda's equation was within ±10% of the experimental values.  相似文献   

5.
Two biofilters fed toluene-polluted air were inoculated with new fungal isolates of either Exophiala oligosperma or Paecilomyces variotii, while a third bioreactor was inoculated with a defined consortium composed of both fungi and a co-culture of a Pseudomonas strain and a Bacillus strain. Elimination capacities of 77 g m–3 h–1 and 55 g m–3 h–1 were reached in the fungal biofilters (with removal efficiencies exceeding 99%) in the case of, respectively, E. oligosperma and Paecilomyces variotii when feeding air with a relative humidity (RH) of 85%. The inoculated fungal strains remained the single dominant populations throughout the experiment. Conversely, in the biofilter inoculated with the bacterial–fungal consortium, the bacteria were gradually overgrown by the fungi, reaching a maximum elimination capacity around 77 g m–3 h–1. Determination of carbon dioxide concentrations both in batch assays and in biofiltration studies suggested the near complete mineralization of toluene. The non-linear toluene removal along the height of the biofilters resulted in local elimination capacities of up to 170 g m–3 h–1 and 94 g m–3 h–1 in the reactors inoculated, respectively, with E. oligosperma and P. variotii. Further studies with the most efficient strain, E. oligosperma, showed that the performance was highly dependent on the RH of the air and the pH of the nutrient solution. At a constant 85% RH, the maximum elimination capacity either dropped to 48.7 g m–3 h–1 or increased to 95.6 g m–3 h–1, respectively, when modifying the pH of the nutrient solution from 5.9 to either 4.5 or 7.5. The optimal conditions were 100% RH and pH 7.5, which allowed a maximum elimination capacity of 164.4 g m–3 h–1 under steady-state conditions, with near-complete toluene degradation.  相似文献   

6.
Summary The ethanol yield was not affected and the ethanol productivity was increased when exponentially decreasing feeding rates were used instead of constant feeding rates in fed batch ethanol fermentations. The influences of the initial sugar feeding rate on the ethanol productivity, on the constant ethanol production rate during the feeding phase and on the initial ethanol production specific rate are represented by Monod-like equations.Nomenclature F reactor feeding rate (L.h–1) - Fo initial reactor feeding rate (L.h–1) - K time constant; see equation (l) (h–1) - ME mass of ethanol in the fermentor (g) - Ms mass of TRS in the fermentor (g) - Mx mass of yeast cells (dry matter) in the fermentor (g) - P ethanol productivity (g.L–1.h–1) - R ethanol constant production rate during the feeding phase (g.h–1) - s standard deviation - So TRS concentration in the feeding mash (g.L–1) - t time (h) - T fermentor filling-up-time (h) - T time necessary to complete the fermentation (h) - TRS total reducing sugars calculated as glucose (g.L–1) - Vo volume of the inoculum (L) - Vf final volume of medium in the fermentor (L) - Xo yeast concentration of the inoculum (dry matter) (g.L–1) - ethanol yield (% of the theoretical value) - initial specific rate of ethanol production (h–1)  相似文献   

7.
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   

8.
The purpose of this work was to investigate the anaerobic transformation ofo-xylene in a laboratory biofilm system with nitrate as an electron acceptor.o-Xylene was degraded cometabolically with toluene as primary carbon source. A mass balance showed thato-xylene was not mineralized but transformed.o-Methyl-benzalcohol ando-methyl-benzaldehyde were identified as intermediates ofo-xylene transformation which resulted in the formation ofo-methyl-benzoic acid as an end product. A cross inhibition phenomenon was observed between toluene ando-xylene. The presence of toluene was necessary for stimulation ofo-xylene transformation, but above a toluene concentration of 1–3 mg/L theo-xylene removal rate dramatically decreased. In returno-xylene inhibited the toluene degradation at concentrations above 2–3 mg/L.  相似文献   

9.
The work reported concerns the removal of mixtures of two ketones, methyl ethyl ketone (MEK) and methyl isobutyl ketone (MIBK), which find wide application as industrial solvents, from effluent air streams in downward flow biofilters operating at relative humidities in excess of 95 percent. The inlet concentrations of the two pollutants were 300 mg m–3 MEK and 330 mg m–3 MIBK. Maximum elimination capacities achieved were 50 g m–3h–1 for MEK and 20 g m–3h–1 for MIBK. Marked interaction between the elimination of the two ketones was observed and established biophysical models for the kinetic analysis of biofilter operation proved inadequate as far as the complex processes involved in multi-component biodegradable vapour elimination were concerned. The complexity of such systems requires further definition and the development of appropriate models for process evaluation and design.  相似文献   

10.
Summary The ethanol yield was not affected and the ethanol productivity increased (10%) when linearly decreasing feeding rates were used instead of constant feeding rates in fed-batch ethanol fermentations.Nomenclature F reactor feeding rate (L.h–1) - ME mass of ethanol in the fermentor (g) - Ms mass of TRS in the fermentor (g) - Mx mass of yeast cells (dry matter) in the fermentor (g) - P ethanol productivity (g.L–1.h–1) - s standard deviation - So TRS concentration in the feeding mash (g.L–1) - t time (h) - T fermentor filling-up time (h) - TRS total reducing sugars calculated as glucose (g.L–1) - Xo yeast cells concentration (dry matter) in the inoculum (g.L–1) - average ethanol yield (% of the theoretical value)  相似文献   

11.
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   

12.
Summary Phenol degradation by free and immobilized cells ofFusarium flocciferum was studied in a chemostat at steady-state conditions. For the free cell system the dilution rates varied from 0.02 to 0.13h–1, with a total phenol removal up to 0.08h–1. Wash-out seemed to set in at 0.11h–1. The immobilized cells showed virtually complete phenol utilization at 1g/l, over a period of four months. At D=0.2h–1 and above 1g/l phenol, the complete phenol removal is not achieved: a progressive increase in the outlet concentration was observed attaining a value of 284mg/l at 1.5g/l.  相似文献   

13.
The kinetics of continuous l-sorbose fermentation using Acetobacter suboxydans with and without cell recycle (100%) were investigated at dilution rates (D) of 0.05, 0.10, 0.15 and 0.3 h–1. The biomass and sorbose concentrations for continuous fermentation without recycle increased as the dilution rate was increased from 0.05 to 0.10 h–1. A maximum biomass concentration of 8.44 g l–1 and sorbose concentration of 176.90 g l–1 were obtained at D=0.10 h–1. The specific rate of sorbose production and volumetric sorbose productivity at this dilution rate were 2.09 g g–1 h–1 and 17.69 g l–1 h–1. However, on further increasing the dilution rate to 0.3 h–1, both biomass and sorbose concentrations decreased to 2.93 and 73.20 g l–1 respectively, mainly due to washout of the reactor contents. However, the specific rate of sorbose formation and volumetric sorbose productivity at this dilution rate increased to 7.49 g g–1 h–1 and 21.96 g l–1 h–1 respectively. Continuous fermentation with 100% cell recycle served to further enhance the concentration of biomass and sorbose to 28.27 and 184.32 g l–1 respectively (in the reactor at a dilution rate of 0.05 h–1). Even though, there was a decline in the biomass and sorbose concentrations to 6.8 and 83.40 g l–1 at a dilution rate of 0.3 h–1, the specific rates of sorbose formation and volumetric sorbose productivity increased to 3.67 g g–1h–1 and 25.02 g l–1 h–1.  相似文献   

14.
Both conventional and genetic engineering techniques can significantly improve the performance of animal cell cultures for the large-scale production of pharmaceutical products. In this paper, the effect of such techniques on cell yield and antibody production of two NS0 cell lines is presented. On the one hand, the effect of fed-batch cultivation using dialysis is compared to cultivation without dialysis. Maximum cell density could be increased by a factor of ~5–7 by dialysis fed-batch cultivation. On the other hand, suppression of apoptosis in the NS0 cell line 6A1 bcl-2 resulted in a prolonged growth phase and a higher viability and maximum cell density in fed-batch cultivation in contrast to the control cell line 6A1 (100)3. These factors resulted in more product formation (by a factor ~2). Finally, the adaptive model-based OLFO controller, developed as a general tool for cell culture fed-batch processes, was able to control the fed-batch and dialysis fed-batch cultivations of both cell lines.Abbreviations A membrane area (dm2) - c Glc,F glucose concentration in nutrient feed (mmol L–1) - c Glc,FD glucose concentration in dialysis feed (mmol L–1) - c Glc,i glucose concentration in inner reactor chamber (mmol L–1) - c Glc,o glucose concentration in outer reactor chamber (dialysis chamber) (mmol L–1) - c Lac,FD lactate concentration in dialysis feed (mmol L–1) - c Lac,i lactate concentration in inner reactor chamber (mmol L–1) - c Lac,o lactate concentration in outer reactor chamber (dialysis chamber) (mmol L–1) - c LS,FD limiting substrate concentration in dialysis feed (mmol L–1) - c LS,i limiting substrate concentration in inner reactor chamber (mmol L–1) - c LS,o limiting substrate concentration in outer reactor chamber (dialysis chamber) (mmol L–1) - c Mab monoclonal antibody concentration (mg L–1) - F D feed rate of dialysis feed (L h–1) - F Glc feed rate of nutrient concentrate feed (L h–1) - K d maximum death constant (h–1) - k d,LS death rate constant for limiting substrate (mmol L–1) - k Glc monod kinetic constant for glucose uptake (mmol L–1) - k Lac monod kinetic constant for lactate uptake (mmol L–1) - k LS monod kinetic constant for limiting substrate uptake (mmol L–1) - K Lys cell lysis constant (h–1) - K S,Glc monod kinetic constant for glucose (mmol L–1) - K S,LS monod kinetic constant for limiting substrate (mmol L–1) - µ cell-specific growth rate (h–1) - µ d cell-specific death rate (h–1) - µ d,min minimum cell-specific death rate (h–1) - µ max maximum cell-specific growth rate (h–1) - P Glc membrane permeation coefficient for glucose (dm h–1) - P Lac membrane permeation coefficient for lactate (dm h–1) - P LS membrane permeation coefficient for limiting substrate (dm h–1) - q Glc cell-specific glucose uptake rate (mmol cell–1 h–1) - q Glc,max maximum cell-specific glucose uptake rate (mmol cell–1 h–1) - q Lac cell-specific lactate uptake/production rate (mmol cell–1 h–1) - q Lac,max maximum cell-specific lactate uptake rate (mmol cell–1 h–1) - q LS cell-specific limiting substrate uptake rate (mmol cell–1 h–1) - q LS,max maximum cell-specific limiting substrate uptake rate (mmol cell –1 h–1) - q Mab cell-specific antibody production rate (mg cell–1 h–1) - q MAb,max maximum cell-specific antibody production rate (mg cell–1 h–1) - t time (h) - V i volume of inner reactor chamber (culture chamber) (L) - V o volume of outer reactor chamber (dialysis chamber) (L) - X t total cell concentration (cells L–1) - X viable cell concentration (cells L–1) - Y Lac/Glc kinetic production constant (stoichiometric ratio of lactate production and glucose uptake) (–)  相似文献   

15.
Chlorella sorokiniana was cultured in heterotrophic or mixotrophic mode in outdoor enclosed tubular photobioreactor. The culture temperature was maintained at 32–35 °C. At night, theChlorella culture grew heterotrophically, and 0.1 M glucose was completely consumed. The biomass growth yield of glucose was 0.35 ± 0.001 g-biomass g-glucose–1. During the day, the algal culture grew mixotrophically and the biomass growth yield was 0.49 g-biomass g-glucose–1 in low density culture (initial biomass concentration, Xo = 2 g l–1), 0.56 g-biomass g-glucose–1 in medium density culture (Xo = 4 g l–1) and 0.46 g-biomass g-glucose–1 in high density culture (Xo = 7 g l–1). The daily area productivity of the culture, with Xo = 4 g l–1 corresponded to 127 g-biomass m–2 d–1 during the day and 79 g-biomass m–2 d–1 during the night. In all the cultures, the dissolved O2 concentration increased in the morning, reached the maximum value at noon, and then decreased in the afternoon. The dissolved CO2 concentration remained at 3 mBar in the morning and increased in the afternoon. Glycolate was not found to accumulate in culture medium.  相似文献   

16.
The biomass and primary production of phytoplankton in Lake Awasa, Ethiopia was measured over a 14 month period, November 1983 to March 1985. The lake had a mean phytoplankton biomass of 34 mg chl a m–3 (n = 14). The seasonal variation in phytoplankton biomass of the euphotic zone (mg chl a m–2 h–1) was muted with a CV (standard deviation/mean) of 31%. The vertical distribution of photosynthetic activity was of a typical pattern for phytoplankton with light inhibition on all but overcast days. The maximum specific rates of photosynthesis or photosynthetic capacity (Ømax) for the lake approached 19 mg O2 (mg chl a)–1 h–1, with high values during periods of low phytoplankton biomass. Areal rates of photosynthesis ranged between 0.30 to 0.73 g O2 m–2 h–1 and 3.3 to 7.8 g O2 m–2 d–1. The efficiency of utilisation of PhAR incident on the lake surface varied from 2.4 to 4.1 mmol E–1 with the highest efficiency observed corresponding to the lowest surface radiation. Calculated on a caloric basis, the efficiency ranged between 1.7 and 2.9%. The temporal pattern of primary production by phytoplankton showed limited variability (CV = 21 %).  相似文献   

17.
Summary Optimal growth conditions for Zymomonas mobilis have been established using continuous cultivation methods. Optimal substrate utilization efficiency occurs with 2.5 g l–1 yeast extract, 2.0 g l–1 ammonium sulfate and 6.0 g l–1 magnesium sulfate in the media. Catabolic activity is at its maximum with glucose uptake rates of 16–18 g l–1 h–1 and ethanol production rates of 8–9 g l–1 h–1, Qg values of 22–26 and Qp values between 11 and 13, which results in 40 g l–1 h–1 ethanol yields using a 100 g l–1 substrate feed. Any increase in these parameters goes on cost of substrate utilization efficiency. Calcium pantothenate can not substitute yeast extract.Abbreviations G Glucose (%) - Pant Calcium pantothenate (mg l–1) - D Dilution rate (h–1) - NH4 Ammonium sulfate (%) - Mg Magnesium sulfate (%) - S1 Residual glucose in the fermenter (g l–1) - S0 Glucose feed (g l–1) - Eth Ethanol concentration (g l–1) - GUR Glucose uptake rate (g l–1 h–1) - Qg Specific glucose uptake rate (g g–1 h–1) - Qp Specific ethanol production rate (g g–1 h–1) - EPR Ethanol production rate (g l–1 h–1) - Yg Yield coefficient for glucose (g g–1) - Yp Conversion efficiency (%) - C Biomass concentration (g l–1) Present address: (Until June 1982) Institut für Mikrobiologie, TH Darmstadt, 6100 Darmstdt, Federal Republic of Germany  相似文献   

18.
A trenching method was used to determine the contribution of root respiration to soil respiration. Soil respiration rates in a trenched plot (R trench) and in a control plot (R control) were measured from May 2000 to September 2001 by using an open-flow gas exchange system with an infrared gas analyser. The decomposition rate of dead roots (R D) was estimated by using a root-bag method to correct the soil respiration measured from the trenched plots for the additional decaying root biomass. The soil respiration rates in the control plot increased from May (240–320 mg CO2 m–2 h–1) to August (840–1150 mg CO2 m–2 h–1) and then decreased during autumn (200–650 mg CO2 m–2 h–1). The soil respiration rates in the trenched plot showed a similar pattern of seasonal change, but the rates were lower than in the control plot except during the 2 months following the trenching. Root respiration rate (R r) and heterotrophic respiration rate (R h) were estimated from R control, R trench, and R D. We estimated that the contribution of R r to total soil respiration in the growing season ranged from 27 to 71%. There was a significant relationship between R h and soil temperature, whereas R r had no significant correlation with soil temperature. The results suggest that the factors controlling the seasonal change of respiration differ between the two components of soil respiration, R r and R h.  相似文献   

19.
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   

20.
Summary Candida tropicalis S001 was grown on the lipid fraction of a protein-containing waste-water in order to (i) remove fat from the water, and (ii) produre yeast biomass for feed. The yeast cells were separated from the waste-water by sedimentation. Defatted waste-water was used for methane production and gave a yield of a 0.3 m3 methane/kg reduced chemical oxygen demand. The maximum specific growth rate (µmax) of C. tropicalis growing on waste-water fat at pH 4.0 was 0.35 h–1; the fat content was decreased from 8 g/l to about 0.1 g/l within 24 h. In continous culture a corresponding reduction was maintained at dilution rates up to 0.36 h–1. The effect on growth of pH, temperature and CO2 concentration was studied with triolein as the major carbon source. The µmax was nearly constant (0.16 h–1) in the pH and temperature range of 3.2–4.0 and 30°–38° C, respectively; 10% CO2 was optimal for growth. Growth on triolein resulted in a biomass yield of 0.70 g dry weight/g fat. Offprint requests to: S. Rydin  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