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.
Structured models of antibiotic fermentation that quantify maturation and aging of product forming biomass are fitted to experimental data. Conditions of superiority of repeated fed batch cultivation are characterized on the basis of a performance criterion that includes penicillin productivity and costs of operation. Emphasis is placed on the relevance of such research to the model aided design of optimal cyclic operation.List of Symbols c IU/mg cost factor - D s–1 dilution rate - J IU · cm–3 · h–1 net productivity - k p IU · mg–11 · h–1 specific product formation rate - k pm IU · mg–1 · h–1 maximum specific product formation rate - p IU/cm3 concentration of penicillin - T s final time of fermentation - t s fermentation time - X kg/m3 concentration of biomass dry weight - X 1kg/m3 concentration of young, immature biomass - X 2 kg/m3 concentration of mature product forming biomass - X c kg/m3 biomass concentration of the end of growth phase - X mkg/m3 maximum biomass concentration Greek Letters s–1 specific maturation rate - s–1 specific aging rate - s–1 specific growth rate - m s–1 maximum specific growth rate - p s–1 specific growth rate during the product formation phase - s cycle time - % volume fraction of draw-off Abbreviations CC chemostat culture - RFBC repeated fed batch culture - RBC repeated batch culture  相似文献   

2.
The effects of a 60 min exposure to photosynthetic photon flux densities ranging from 300 to 2200 mol m–2s–1 on the photosynthetic light response curve and on PS II heterogeneity as reflected in chlorophyll a fluorescence were investigated using the unicellular green alga Chlamydomonas reinhardtii. It was established that exposure to high light acts at three different regulatory or inhibitory levels; 1) regulation occurs from 300 to 780 mol m–2s–1 where total amount of PS II centers and the shape of the light response curve is not significantly changed, 2) a first photoinhibitory range above 780 up to 1600 mol m–2s–1 where a progressive inhibition of the quantum yield and the rate of bending (convexity) of the light response curve can be related to the loss of QB-reducing centers and 3) a second photoinhibitory range above 1600 mol m–2s–1 where the rate of light saturated photosynthesis also decreases and convexity reaches zero. This was related to a particularly large decrease in PS II centers and a large increase in spill-over in energy to PS I.Abbreviations Chl chlorophyll - DCMU 3,(3,4-dichlorophenyl)-1,1-dimethylurea - FM maximal fluorescence yield - Fpl intermediate fluorescence yield plateau level - F0 non-variable fluorescence yield - Fv total variable fluorescence yield (FM-F0) - initial slope to the light response curve, used as an estimate of initial quantum yield - convexity (rate of bending) of the light response curve of photosynthesis - LHC light-harvesting complex - Pmax maximum rate of photosynthesis - PQ plastoquinone - Q photosynthetically active photon flux density (400–700 nm, mol m–2s–1) - PS photosystem - QA and QB primary and secondary quinone electron acceptor of PS II  相似文献   

3.
Biochemical and biophysical parameters, including D1-protein turnover, chlorophyll fluorescence, oxygen evolution activity and zeaxanthin formation were measured in the marine seagrassZostera capricorni (Aschers) in response to limiting (100 mol·m–2·–1), saturating (350 mol·m–2·s–1) or photoinhibitory (1100 mol·m–2·s–1) irradiances. Synthesis of D1 was maximal at 350 mol·m–2·s–1 which was also the irradiance at which the rate of photosynthetic O2 evolution was maximal. Degradation of D1 was saturated at 350 mol·m–2·s–1. The rate of D1 synthesis at 1100 mol·m–2·s–1 was very similar to that at 350 mol·m–2·s–1 for the first 90 min but then declined. At limiting or saturating irradiance little change was observed in the ratio of variable to maximal fluorescence (Fv/Fm) measured after dark adaptation of the leaves, while significant photoinhibition occurred at 1100 mol·m–2·s–1. The proportion of zeaxanthin in the total xanthophyll pool increased with increasing irradiance, indicative of the presence of a photoprotective xanthophyll cycle in this seagrass. These results are consistent with a high level of regulatory D1 turnover inZostera under non-photoinhibitory irradiance conditions, as has been found previously for terrestrial plants.We would like to thank Professor Peter Böger (Department of Plant Biochemistry, University of Konstanz, Germany) for the kind gift of D1 antibodies. This work was partly supported by a University of Queensland Enabling Grant to CC.  相似文献   

4.
Various ecophysiological investigations on carnivorous plants in wet soils are presented. Radial oxygen loss from roots of Droseraceae to an anoxic medium was relatively low 0.02 – 0.07 mol(O2) m– 2 s–1 in the apical zone, while values of about one order of magnitude greater were found in both Sarracenia rubra roots and Genlisea violacea traps. Aerobic respiration rates were in the range of 1.6 – 5.6 mol kg–1 (f.m.) s–1 for apical root segments of seven carnivorous plant species and 0.4 – 1.1 mol kg–1 (f.m.) s–1 for Genlisea traps. The rate of anaerobic fermentation in roots of two Drosera species was only 5 – 14 % of the aerobic respiration. Neither 0.2 mM NaN3 nor 0.5 mM KCN influenced respiration rate of roots and traps. In all species, the proportion of cyanide-resistant respiration was high and amounted to 65 – 89 % of the total value. Mean rates of water exudation from excised roots of 12 species ranged between 0.4 – 336 mm 3 kg–1 (f.m.) s–1 with the highest values being found in the Droseraceae. Exudation from roots was insensitive to respiration inhibitors. No significant difference was found between exudation rates from roots growing in situ in anoxic soil and those kept in an aerated aquatic medium. Carnivorous plant roots appear to be physiologically very active and well adapted to endure permanent soil anoxia.  相似文献   

5.
The apparent viscosity of non-Newtonian fermentation media is examined. The present state of this subject is discussed. The energy dissipation rate concept is used for a new evaluation of the apparent viscosity in bioreactors, i.e. stirred tank and bubble column bioreactors. The proposed definition of the apparent viscosity is compared with the definitions available in the literature.List of Symbols A d m 2 downcomer cross-sectional area - A r m 2 riser cross-sectional area - a m–1 specific surface area - C constant in eq. (13) - D m column diameter - D I m impeller diameter - g m s–2 gravitational acceleration - h J m–2 s–1 K–1 heat transfer coefficient - K Pa s n consistency index in a power-law model - k constant in eq. (3) - k L m s –1 liquid-phase mass transfer coefficient - N s–1 impeller speed - n flow index in a power-law model - P W power input - Re Reynolds number ND I /2 /(/) - U sg m s –1 superficial gas velocity - (U sg ) r m s–1 superficial gas velocity based on riser - V-m3 liquid volume - v 0 m s–1 friction velocity Greek Symbols s–1 shear rate - c s–1 characteristic shear rate - W kg–1 energy dissipation rate per unit mass - W kg–1 characteristic energy dissipation rate per unit mass - Pa s viscosity - app Pa s apparent viscosity - kg m–3 density - Pa shear stress  相似文献   

6.
A mathematical model for pellet development of filamentous microorganisms is presented, which simulates in detail location and growth of single hyphal elements. The basic model for growth, septation and branching of discrete hyphae is adopted from Yang et al. [2, 23]. Exact solutions to the intracellular mass-balance equations of a growth-limiting key component is given for two types of either branched or unbranched cellular compartments. Furthermore, the growth model was extended in regard to the external mass-balance equations of limiting substrates (oxygen, glucose) under the assumption that the substrates can enter the denser regions of the pellet only diffusively. Penetration of the substrates into the more porous outer regions of the pellet occurs more easily due to microeddies in the surrounding fluid. Chipping of hyphae from the pellet surface by shear forces was included in the model as well. The application of shear forces leads to a marked smoothing of the simulated pellet surface. The development of pellets from spore germination up to late stages with cell-lysis due to shortage of substrates in the pellet centre can be described. The effects of various model parameters are discussed.List of Symbols A i algebraic coefficient (i = 1, 2,..., 6) - B i algebraic coefficient (i = 1, 2,..., 6) - C i mass-concentration of component i (i = O2, S) (gl–1) - C i,crit concentration of substance i critical for lysis (i=O2, S) (gl–1) - C i,stop concentration of substance i below which cells are inactivated (gl–1) - C(l i,t) intracellular concentration of the key component at site l i and time t (gl–1) - C m maximal intracellular concentration of the key component (gl–1) - C X Concentration of dry biomass (gl–1) - D intracellular diffusion coefficient of the key component (m2 h–1) - D max,i maximal molecular diffusion coefficient of substrate i (i = O2, S) (m2 h–1) - D eff,i effective diffusion coefficient of component i (i = O2) (m2 h–1) - d h cross-sectional diameter of hyphae (m) - k production coefficient for the key component (h–1) - K s Monod coefficient for glucose (gl–1) - k 0 Monod coefficient for oxygen (gl–1) - L c total length of a compartment (m) - L i total length of branch i (i=1, 2, 3) (m) - l i position on branch i (i=1, 2, 3) - L m maximal length of a segment (m) - m i maintenance coefficient of substrate i (h–1) - N m maximal number of segments in a compartment - n iR number of tips of type i in layer R, i=1, 2 - p auxiliary variable (see Eq. (7)) - P Br probability that a hypha is chipped off (%h–1) - pO 2 partial pressure of oxygen in the liquid phase (%) - Q auxiliary variable (see Eq. (8)) - Q i uptake rate of substrate i (i = O2, S) (gl–1 h–1) - q auxiliary variable (see Eq. (7)) - R index of radial layer (R=1, 2, 3,..., R max) - r radius (m) - r crit critical radius, Eq. (15) (m) - r max pellet radius (m) - r tip distance from the pellet centre to the tip position (m) - r thr threshold radius (m) - s auxiliary variable (see Eq. (7)) - S index for glucose - t time (h) - v R volume of layer R (1) - Y Mi observable yield coefficient of biomass on substrate i (gg–1) - Y Xi yield coefficient of biomass on substrate i (gg–1) Greek Letters i actual tip expansion rate (m h–1) - i,m actual maximal extension rate of tip i (i=1, 2) (m h–1) - 1y lysis rate (h–1) - m maximal tip extension rate (m h–1) - auxiliary variable in Eq. (2) - auxiliary variable in Eq. (3) - auxiliary variable defined in Eq. (4) (m–1) - shear shear force parameter - R overall specific growth rate in layer R (h–1) - m maximal specific hyphal growth rate (h–1) - cell volume density (l cell volume per 1) - crit critical cell volume density in Eq. (15) - S shear force parameter - X cell mass density (g dry weight per 1 wet cells) - (C i) growth kinetics on substrate i - proportional factor in Eq. (34) (l g–1) We thank the Deutsche Forschungsgemeinschaft (DFG) for financially supporting parts of this work.We thank the Deutsche Forschungsgemeinschaft (DFG) for financially supporting parts of this work.  相似文献   

7.
The light-dependent rate of photosystem-II (PSII) damage and repair was measured in photoautotrophic cultures of Dunaliella salina Teod. grown at different irradiances in the range 50–3000 mol photons · m–2· s–1. Rates of cell growth increased in the range of 50–800 mol photons·m–2·s–1, remained constant at a maximum in the range of 800–1,500 mol photons·m–2 ·s–1, and declined due to photoinhibition in the range of 1500–3000 mol photons·m–2·s–1. Western blot analyses, upon addition of lincomycin to the cultures, revealed first-order kinetics for the loss of the PSII reaction-center protein (D1) from the 32-kDa position, occurring as a result of photodamage. The rate constant of this 32-kDa protein loss was a linear function of cell growth irradiance. In the presence of lincomycin, loss of the other PSII reaction-center protein (D2) from the 34-kDa position was also observed, occurring with kinetics similar to those of the 32-kDa form of D1. Increasing rates of photodamage as a function of irradiance were accompanied by an increase in the steady-state level of a higher-molecular-weight protein complex ( 160-kDa) that cross-reacted with D1 antibodies. The steady-state level of the 160-kDa complex in thylakoids was also a linear function of cell growth irradiance. These observations suggest that photodamage to D1 converts stoichiometric amounts of D1 and D2 (i.e., the D1/D2 heterodimer) into a 160-kDa complex. This complex may help to stabilize the reaction-center proteins until degradation and replacement of D1 can occur. The results indicated an intrinsic half-time of about 60 min for the repair of individual PSII units, supporting the idea that degradation of D1 after photodamage is the rate-limiting step in the PSII repair process.Abbreviations Chl chlorophyll - PSI photosystem I - PSII photosystem II - D1 the 32-kDa reaction-center protein of PSII, encoded by the chloroplast psbA gene - D2 the 34-kDa reactioncenter protein of PSII, encoded by the chloroplast psbD gene - QA primary electron-accepting plastoquinone of PSII The work was supported by grant 94-37100-7529 from the US Department of Agriculture, National Research Initiative Competitive Grants Program.  相似文献   

8.
To gain information on extended flight energetics, quasi-natural flight conditions imitating steady horizontal flight were set by combining the tetheredflight wind-tunnel method with the exhaustion-flight method. The bees were suspended from a two-component aerodynamic balance at different, near optimum body angle of attack and were allowed to choose their own speed: their body mass and body weight was determined before and after a flight; their speed, lift, wingbeat frequency and total flight time were measured throughout a flight. These values were used to determine thrust, resultant aerodynamic force (magnitude and tilting angle), Reynolds number, total flight distance and total flight impulse. Flights in which lift was body weight were mostly obtained. Bees, flown to complete exhausion, were refed with 5, 10, 15 or 20 l of a 1.28-mol·l-1 glucose solution (energy content w=18.5, 37.0, 55.5 or 74.0 J) and again flown to complete exhaustion at an ambient temperature of 25±1.5°C by a flight of known duration such that the calculation of absolute and relative metabolic power was possible. Mean body mass after exhaustion was 76.49±3.52 mg. During long term flights of 7.47–31.30 min similar changes in flight velocity, lift, thrust, aerodynamic force, wingbeat frequency and tilting angle took place, independent of the volume of feeding solution. After increasing rapidly within 15 s a more or less steady phase of 60–80% of total flight time, showing only a slight decrease, was followed by a steeper, more irregular decrease, finally reaching 0 within 20–30 s. In steady phases lift was nearly equal to resultant aerodynamic force; tilting angle was 79.8±4.0°, thrust to lift radio did not vary, thrust was 18.0±7.4% of lift, lift was somewhat higher/equal/lower than body mass in 61.3%, 16.1%, 22.6% of all totally analysable flights (n=31). The following parameters were varied as functions of volume of feeding solution (5–20 l in steps of 5 l) and energy content. (18.5–74.0 J in steps of 18.5 J): total flight time, velocity, total flight distance, mean lift, thrust, mean resultant aerodynamic force, tilting angle, total flight impulse, wingbeat frequency, metabolic power and metabolic power related to body mass, the latter related to empty, full and mean (=100 mg) body mass. The following positive correlations were found: L=1.069·10-9 f 2.538; R=1.629·10-9 f 2.464; P m=7.079·10-8 f 2.456; P m=0.008v+0.008; P m=18.996L+0.022; P m=19.782R+0.021; P m=82.143T+0.028; P m=1.245·bm f 1.424 ; P mrel e=6.471·bm f 1.040 ; =83.248+0.385. The following negative correlations were found: V=3.939–0.032; T=1.324·10-4–0.038·10-4. Statistically significant correlations were not found in T(f), L(), R(), f(), P m(bm e), P m rel e(bm e), P m rel f(bm e), P m rel f(bm f).Abbreviations A(m2) frontal area - bl(m) body length - bm(mg) body mass - c(mol·1-1) glucose concentration of feeding solution - c D (dimensionless) drag coefficient, related to A - D(N) drag - F w(N) body weight - F wp weight of paper fragment lost at flight start - f wingbeat frequency (s-1) - g(=9.81 m·s-2) gravitational acceleration - I(Ns)=R(t) dt total impulse of a flight - L(N) lift vertical sustaining force component - P m(J·s-1=W) metabolic power - Pm ret (W·g-1) metabolic power, related to body mass - R(N) resultant aerodynamic force - Re v·bl·v -1 (dimensionless) Reynolds number, related to body length - s(m) v(t) dt virtual flight distance of a flight - s(km) total virtual flight distance - T (N) thrust horizontal force component of horizontal flight - T a (°C) ambient temperature - t(s) time - t tot (s or min) total flight time - v(m·s-1) flight velocity - v(l) volume of feeding solution - W (J) energy and energy content of V - ( °) body angle of attack between body longitudinal axis and flow direction - ( °) tilting angle ( 90°) between R and the horizont in horizontal flight v(=1.53·10-5m2·s-1 for air at 25°) kinematic viscosity - (=1.2 kg·m-3 at 25°C) air density  相似文献   

9.
Single leaf photosynthetic characteristics of Alnus glutinosa, A. incana, A. rubra, Elaeagnus angustifolia, and E. umbellata seedlings conditioned to ambient sunlight in a glasshouse were assessed. Light saturation occurred between 930 and 1400 mol m-2s-1 PAR for all species. Maximum rates of net photosynthesis (Pn) measured at 25°C ranged from 12.8 to 17.3 mol CO2m-2s-1 and rates of dark respiration ranged from 0.74 to 0.95 mol CO2m-2s-1. These values of leaf photosynthetic variables are typical of early to midsuccessional species. The rate of Pn measured at optimal temperature (20°C) and 530mol m-2s-1 PAR was significantly (p<0.01) correlated with leaf nitrogen concentration (r=0.69) and negatively correlated with the mean area of a leaf (r=–0.64). We suggest that the high leaf nitrogen concentration and rate of Pn observed for Elaeagnus umbellata and to a lesser degree for E. angustifolia are genetic adaptations related to their crown architecture.Abbreviations Pn net photosynthesis  相似文献   

10.
Plant height, light-saturated rates of photosynthesis (A max) and foliar nitrogen concentration (N 1) were measured forBartsia trixago under field conditions in Mallorca. All three variables were postively correlated, and were also positively related to the abundance of nitrogen-fixing legumes in the associated vegetation (putative host species).A max forB. trixago ranged from 7.7 to 18.8 mol m-2 s-1; similar rates were measured for a second hemiparasiteParentucellia viscosa, and both species were within the range of rates measured for six putative hosts (10.6–19.2 mol m-2 s-1). Fertilization of unattachedB. trixago plants with inorganic nitrogen (ammonium nitrate) elicited neither the growth nor the photosynthetic responses observed in plants considered to be parasitic on legumes and in receipt of an enriched organic nitrogen supply. Both hemiparasites had high diurnal leaf conductances (g s) (469–2291 mmol m-2 s-1) and were at the upper end of the range of those measured in putative hosts (409–879 mmol m-2 s-1). In contrast with the latter, high nocturnal rates ofg s were also recorded for the two hemiparasites (517–1862 mmol m-2 s-1). There was no clear relationship between eitherA max orN 1 and eitherg s, transpiration (E) or water use efficiency (A max/E) inB. trixago plants. The economics of water loss appear to be independent of both the supply of nitrogen from the host and autotrophic carbon fixation.  相似文献   

11.
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   

12.
We have isolated Chl a-Chl c-carotenoid binding proteins from the dinoflagellates Prorocentrum minimum and Heterocapsa pygmaea grown under high (500 mol m–2 s–1, HL) and low (35 mol m–2 s–1, LL) light conditions. We compared various isolation procedures of membrane bound light harvesting complexes (LHCs) and assayed the functionality of the solubilized proteins by determining the energy transfer efficiency from the accessory pigments to Chl a by means of fluorescence excitation spectra. The identity of the newly isolated protein-complexes were confirmed by immunological cross-reactions with antibodies raised against the previously described membrane bound Chl a-c proteins (Boczar et al. (1980) FEBS Lett 120: 243–247). Spectroscopic analysis demonstrated the relatedness of these proteins with the recently described Chl-a-c 2-peridinin (ACP) binding protein (Hiller et al. (1993) Photochem Photobiol 57: 125–131; Iglesias Prieto et al. (1993) Phil Trans R Soc London B 338: 381–392). The water-soluble peridinin-Chl a binding-protein (PCP) was not detectable in P. minimum. Two functional forms of ACP with different pigmentation were isolated. A variant of ACP which was isolated from high-light grown cells, that specifically binds increased amounts of diadinoxanthin was compared to the previously described ACPs that bind proportionately more peridinin.Abbreviations ACP Chl a-Chl c-peridinin binding protein - AEBSF 4-(2-aminoethyl)-benzenesulfonyl fluoride hydrochloride - DDM dodecyl -d maltoside - Deriphat 160 N-lauryl-beta-iminopropionic acid - HEPES (N-2-hydroxyethylpiparizine-N-2-ethanesulphonic acid) - HL high light (500 mol m–2 s–1) - LL low light (35 mol m–2 s–1) - 730 fluorescence yield (emission at 730 nm) - PCP peridinin-Chl a-binding protein - PMSF phenyl-methyl-sulfonyl-fluoride - PS I Photosystem I - PS II Photosystem II  相似文献   

13.
Summary Reaction kinetic analysis of the electrical properties of the electrogenic Cl pump inAcetabularia has been extended from steady-state to nonsteady-state conditions: electrical frequency responses of theAcetabularia membrane have been measured over the range from 1 Hz to 10 kHz at transmembrane potential differences across the plasmalemma (V m ) between –70 and –240 mV using voltage-clamp techniques. The results are well described by an electrical equivalent circuit with three parallel limbs: a conventional membrane capacitancec m , a steadystate conductanceg o (predominantly of the pump pathway plus a minor passive ion conductance) and a conductanceg s in series with a capacitancec p which are peculiar to the temporal behavior of the pump. The absolute values and voltage sensitivities of these four elements have been determined:c m of about 8 mF m–2 turned out to be voltage insensitive; it is considered to be normal.g o is voltage sensitive and displays a peak of about 80 S m–2 around –180 mV. Voltage sensitivity ofg s could not be documented due to large scatter ofg s (around 80 S m–2).c p behaved voltage sensitive with a notch of about 20 mF m–2 around –180 mV, a peak of about 40 mF m–2 at –120 mV and vanishing at –70 mV. When these data are compared with the predictions of nonsteady-state electrical properties of charge transport systems (U.-P. Hansen, J. Tittor, D. Gradmann, 1983,J. Membrane Biol. in press), model A (redistribution of states within the reaction cycle) consistently provides magnitude and voltage sensitivity of the elementsg o ,g s andc p of the equivalent circuit, when known kinetic parameters of the pump are used for the calculations. This analysis results in a density of pump elements in theAcetabularia plasmalemma of about 50 nmol m–2. The dominating rate constants for the redistribution of the individual states of the pump in the electric field turn out to be in the range of 500 sec–1, under normal conditions.  相似文献   

14.
Effects of light and temperature, on the growth of three freshwater green algae isolated from an eutrophic lake and identified as Selenastrum minutum, Coelastrum microporum f. astroidea and Cosmarium subprotumidumwere studied in batch cultures under non-nutrient limited conditions. Experiments were performed to determine the growth rate over a wide range of light intensities (30–456 mol m–2 s–1) and temperature (15–35°C), using a 15/9 (light/dark) photoperiod cycle. The maximum growth rates and the optimum light intensities at a temperature of 35°C were 1.73 d–1 and 420 mol m–2 s–1for Selenastrum minutum, 1.64 d–1 and 400 mol m–2 s–1 for Coelastrum microporum and 1.00 d–1 and 400 mol m–2 s1 for Cosmarium subprotumidum. The results were fitted with the mathematical models of Steele (1965), Platt & Jassby (1976) and Peeters & Eilers (1978). Steele's function and equation of Platt & Jassby don't describe correctly the relationship between the growth and light intensity. In the opposite, the equation of Peeters & Eilers provides the best fit for the three species.  相似文献   

15.
Saccharomyces cerevisiae-based ethanol fermentations were conducted in batch culture, in a single stage continuous stirred tank reactor (CSTR), a multistage CSTR, and in a fermentor contaminated with Lactobacillus that corresponded to the first fermentor of the multistage CSTR system. Using a glucose concentration of 260 g l–1 in the medium, the highest ethanol concentration reached was in batch (116gl–1), followed by the multistage CSTR (106gl–1), and the single stage CSTR continuous production system (60gl–1). The highest ethanol productivity at this sugar concentration was achieved in the multistage CSTR system where a productivity of 12.7gl–1h–1 was seen. The other fermentation systems in comparison did not exceed an ethanol productivity of 3gl–1h–1. By performing a continuous ethanol fermentation in multiple stages (having a total equivalent working volume of the tested single stage), a 4-fold higher ethanol productivity was achieved as compared to either the single stage CSTR, or the batch fermentation.  相似文献   

16.
In Taxus cuspidata callus, vanadyl sulfate (10 mg l–1) induced a high (146 g g–1 dry wt) production of 10-deacetylbaccatin III in comparison to 7 g g–1 dry wt of the control. The content of paclitaxel in this species increased from 16 g g–1 to 74 g g–1 dry wt when 20 mg phenylalanine l–1 was used. In T. media, p-aminobenzoic acid induced the highest content of 10-deacetylbaccatin III (481 g g–1 dry wt) versus 181 g g–1 in the control. Paclitaxel increased from 89 to 139 g g–1 dry wt after adding chitosan (20 mg l–1) to the cultures.  相似文献   

17.
The light-dependent modulation of ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) activity was studied in two species: Phaseolus vulgaris L., which has high levels of the inhibitor of Rubisco activity, carboxyarabinitol 1-phosphate (CA1P), in the dark, and Chenopodium album L., which has little CA1P. In both species, the ratio of initial to fully-activated Rubisco activity declined by 40–50% within 60 min of a reduction in light from high a photosynthetic photon flux density (PPFD; >700 mol · m–2 · s–1) to a low PPFD (65 ± 15 mol · m–2 · s–1) or to darkness, indicating that decarbamylation of Rubisco is substantially involved in the initial regulatory response of Rubisco to a reduction in PPFD, even in species with potentially extensive CA1P inhibition. Total Rubisco activity was unaffected by PPFD in C. album, and prolonged exposure (2–6 h) to low light or darkness was accompanied by a slow decline in the activity ratio of this species. This indicates that the carbamylation state of Rubisco from C. album gradually declines for hours after the large initial drop in the first 60 min following light reduction. In P. vulgaris, the total activity of Rubisco declined by 10–30% within 1 h after a reduction in PPFD to below 100 mol · m–2 · s–1, indicating CA1P-binding contributes significantly to the reduction of Rubisco capacity during this period, but to a lesser extent than decarbamylation. With continued exposure of P. vulgaris leaves to very low PPFDs (< 30 mol · m–2 · s–1), the total activity of Rubisco declined steadily so that after 6–6.5 h of exposure to very low light or darkness, it was only 10–20% of the high-light value. These results indicate that while decarbamylation is more prominent in the initial regulatory response of Rubisco to a reduction in PPFD in P. vulgaris, binding of CA1P increases over time and after a few hours dominates the regulation of Rubisco activity in darkness and at very low PPFDs.Abbreviations CA1P 2-carboxyarabinitol 1-phosphate - CABP 2-carboxyarabinitol 1,5-bisphosphate - kcat substrate-saturated turnover rate of fully carbamylated enzyme - PPFD photosynthetically active photon flux density (400–700 nm) - Rubisco ribulose-1,5-bisphosphate carboxylase/oxygenase - RuBP ribulose-1,5-bisphosphate  相似文献   

18.
Pinus pumila (Pallas) Regel. is a dominant dwarf tree in alpine regions of Japan. The possible factors limiting the net photosynthetic rate (Pn) of the needles of P. pumila were examined in the snow-melting (May) and the summer (August) seasons. In August, in situ maximum Pn was 20 mol kg–1 needle s–1 in the current-year needles and 25 mol kg–1 needle s–1 in the 1-year-old needles. Diurnal trends of Pn in August were positively related to fluctuations in photosynthetic photon flux density (PPFD) and no midday depression of Pn was found, indicating that a decrease in PPFD rather than an increase in needle-to-air vapor pressure deficit (W) might cause the reduction of Pn. Both stomatal conductance (gs) and Pn were lower in May than in August. In May, Pn and gs were almost zero in the morning, but gradually increased with decreasing W, reaching maximum Pn values (4 mol kg–1 needle s–1) and gs (60 mmol kg–1 needle s–1) at 16.00 hours. The daytime Pn in May was positively related to gs. Relative water content in the exposed needles above the snow in May was 83%, which was far above the lethal level. This indicates that the water flow from stems or soils to needles was enough to compensate for a small amount of water loss due to the low gs in May, although the water supplied to needles would be impeded by the low temperatures. Thus, the reduced gs in May would be important for avoiding needle desiccation, and would result in a reduced Pn.  相似文献   

19.
A thermophilic bacterium, which we designated as Geobacillus thermoleovorans 47b was isolated from a hot spring in Beppu, Oita Prefecture, Japan, on the basis of its ability to grow on bitter peptides as a sole carbon and nitrogen source. The cell-free extract from G. thermoleovorans 47b contained leucine aminopeptidase (LAP; EC 3.4.11.10), which was purified 164-fold to homogeneity in seven steps, using ammonium sulfate fractionation followed by the column chromatography using DEAE-Toyopearl, hydroxyapatite, MonoQ and Superdex 200 PC gel filtration, followed again by MonoQ and hydroxyapatite. The enzyme was a single polypeptide with a molecular mass of 42,977.2 Da, as determined by matrix-assisted laser desorption ionization and time-of-flight mass spectrometry, and was found to be thermostable at 90°C for up to 1 h. Its optimal pH and temperature were observed to be 7.6–7.8 and 60°C, respectively, and it had high activity towards the substrates Leu-p-nitroanilide (p-NA)(100%), Arg-p-NA (56.3%) and LeuGlyGly (486%). The Km and Vmax values for Leu-p-NA and LeuGlyGly were 0.658 mM and 25.0 mM and 236.2 mol min–1 mg–1 protein and 1,149 mol min–1 mg–1 protein, respectively. The turnover rate (kcat) and catalytic efficiency (kcat/ Km) for Leu-p-NA and LeuGlyGly were 10,179 s–1 and 49,543 s–1 and 15,470 mM–1 s–1 and 1981.7 mM–1 s–1, respectively. The enzyme was strongly inhibited by EDTA, 1,10-phenanthroline, dithiothreitol, -mercaptoethanol, iodoacetate and bestatin; and its apoenzyme was found to be reactivated by Co2+ .  相似文献   

20.
Young sporophytes of short-stipe ecotype ofEcklonia cavafrom a warmer locality (Tei, Kochi Pref., southern Japan) and those of long-stipe ecotype from a cooler locality (Nabeta, Shizuoka Pref., central Japan) were transplanted in 1995 to artificial reefs immersed at the habitat of long-stipe ecotype in Nabeta Bay, Shizuoka Pref., central Japan. The characteristics of photosynthesis and respiration of bladelets of the transplanted sporophytes of the two ecotypes were compared in winter and summer 1997; the results were assessed per unit area, per unit chlorophyllacontent and per unit dry weight. In photosynthesis-light curves at 10–29 °C, light saturation occurred at 200–400 mol photon m–2s–1in sporophytes from both Tei and Nabeta. The maximum photosynthetic rate (P max) at 10–29 °C and the light-saturation index (I k) at 25–29 °C in sporophytes from both localities were generally higher in winter than in summer.P maxat 25–29 °C (per unit area and chlorophylla) were higher in sporophytes from Tei than those from Nabeta in both seasons. The optimum temperature for photosynthesis was 25 °C in winter and 27 °C in summer at high light intensities of 100–400 mol photon m–2s–1. However, at lower light intensities of 12.5–50 mol photon m–2s–1, it was 20 °C in winter and 25–27 °C in summer for sporophytes from both locations. Dark respiration increased with temperature rise in the range of 10–29 °C in sporophytes from both locations in summer and winter. The sporophytes transplanted from Tei (warmer area) showed higher photosynthetic activities than those from Nabeta (cooler area) at warmer temperatures even under the same environmental conditions. This indicates that these physiological ecotypes have arisen from genetic differentiation.  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