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.
Limiting nutrient patchiness is examined as a factor affecting the community structure and species succession of natural phytoplankton communities held in ammonia limited continuous culture at a dilution rate of 0.3 day?1. It was found that under a homogeneous distribution of the limiting nutrient members of genus Chaetoceros dominated and when ammonia was added daily (patchy distribution), Skeletonema dominated. Intermediate patchiness gave rise to an assemblage dominated by both Chaetoceros and Skeletonema. The nutrient uptake ability of each assemblage was determined three weeks after experiment initiation. Each assemblage was best able to optimize uptake of ammonia under its particular patchy nutrient regime. Optimization of a patchy environment took place by an increased maximal uptake rate (Vmax) while optimization of a homogeneous environment appeared to take place by increased substrate affinity (i.e., low Ks).This study demonstrates that limiting nutrient patchiness can alter the relative abundance of populations within a community based on each population's ability to exploit the limiting resource under a particular degree of patchiness. We also show that coexistence of two populations might be expected due to the patchiness of a single limiting nutrient. The importance of patchiness in relation to other factors which determine community structure is discussed.  相似文献   

2.
Marine phytoplankton and macroalgae acquire important resources, such as inorganic nitrogen, from the surrounding seawater by uptake across their entire surface area. Rates of ammonium and nitrate uptake per unit surface area were remarkably similar for both marine phytoplankton and macroalgae at low external concentrations. At an external concentration of 1 μM, the mean rate of nitrogen uptake was 10±2 nmol·cm?2·h?1 (n=36). There was a strong negative relationship between log surface area:volume (SA:V) quotient and log nitrogen content per cm2 of surface (slope=?0.77), but a positive relationship between log SA:V and log maximum specific growth rate (μmax; slope=0.46). There was a strong negative relationship between log SA:V and log measured rate of ammonium assimilation per cm2 of surface, but the slope (?0.49) was steeper than that required to sustain μmax (?0.31). Calculated rates of ammonium assimilation required to sustain growth rates measured in natural populations were similar for both marine phytoplankton and macroalgae with an overall mean of 6.2±1.4 nmol·cm?2·h?1 (n=15). These values were similar to maximum rates of ammonium assimilation in phytoplankton with high SA:V, but the values for algae with low SA:V were substantially less than the maximum rate of ammonium assimilation. This suggests that the growth rates of both marine phytoplankton and macroalgae in nature are often constrained by rates of uptake and assimilation of nutrients per cm2 surface area.  相似文献   

3.
Emiliania huxleyi (strain L) expressed an exceptional P assimilation capability. Under P limitation, the minimum cell P content was 2.6 fmol P·cell?1, and cell N remained constant at all growth rates at 100 fmol N·cell?1. Both, calcification of cells and the induction of the phosphate uptake system were inversely correlated with growth rate. The highest (cellular P based) maximum phosphate uptake rate (VmaxP) was 1400 times (i.e. 8.9 h?1) higher than the actual uptake rate. The affinity of the P‐uptake system (dV/dS) was 19.8 L·μmol?1·h?1 at μ = 0.14 d?1. This is the highest value ever reported for a phytoplankton species. Vmax and dV/dS for phosphate uptake were 48% and 15% lower in the dark than in the light at the lowest growth rates. The half‐saturation constant for growth was 1.1 nM. The coefficient for luxury phosphate uptake (Qmaxt/Qmin) was 31. Under P limitation, E. huxleyi expressed two different types of alkaline phosphatase (APase) enzyme kinetics. One type was synthesized constitutively and possessed a Vmax and half‐saturation constant of 43 fmol MFP·cell?1·h?1 and 1.9 μM, respectively. The other, inducible type of APase expressed its highest activity at the lowest growth rates, with a Vmax and half‐saturation constant of 190 fmol MFP·cell?1·h?1 and 12.2 μM, respectively. Both APase systems were located in a lipid membrane close to the cell wall. Under N‐limiting growth conditions, the minimum N quotum was 43 fmol N·cell?1. The highest value for the cell N‐specific maximum nitrate uptake rate (VmaxN) was 0.075 h?1; for the affinity of nitrate uptake, 0.37 L·μmol?1·h?1. The uptake rate of nitrate in the dark was 70% lower than in the light. N‐limited cells were smaller than P‐limited cells and contained 50% less organic and inorganic carbon. In comparison with other algae, E. huxleyi is a poor competitor for nitrate under N limitation. As a consequence of its high affinity for inorganic phosphate, and the presence of two different types of APase in terms of kinetics, E. huxleyi is expected to perform well in P‐controlled ecosystems.  相似文献   

4.
In the early nineties, Undaria pinnatifida has been accidentally introduced to Nuevo Gulf (Patagonia, Argentina) where the environmental conditions would have favored its expansion. The effect of the secondary treated sewage discharge from Puerto Madryn city into Nueva Bay (located in the western extreme of Nuevo Gulf) is one of the probable factors to be taken into account. Laboratory cultures of this macroalgae were conducted in seawater enriched with the effluent. The nutrients (ammonium, nitrate and phosphate) uptake kinetics was studied at constant temperature and radiation (16?°C and 50 μE m?2 s?1 respectively). Uptake kinetics of both inorganic forms of nitrogen were described by the Michaelis–Menten model during the surge phase (ammonium: V max sur: 218.1 μmol h?1 g?1, K s sur: 476.5 μM and nitrate V max sur: 10.7 μmol h?1 g?1, K s sur: 6.1 μM) and during the assimilation phase (ammonium: V max ass: 135.6 μmol h?1 g?1, K s ass: 407.2 μM and nitrate V max ass: 1.9 μmol h?1 g?1, K s ass: 2.2 μM), with ammonium rates always higher than those of nitrate. Even though a net phosphate disappearance was observed in all treatments, uptake kinetics of this ion could not be properly estimated by the employed methodology.  相似文献   

5.
The brown seaweeds Fucus distichus Linnaeus subsp. edentatus (de la Pylaie) Powell, a perennial, and Chordaria flagelliformis (O.F. Mueller) C. Agardh, a summer annual, were grown in continuous-flow greenhouse cultures. In winter, the growth of Fucus distichus ( μ = 0.01 · day?1 ) in culture was not stimulated by added ammonium. Nitrogen in excess of the growth requirement was accumulated in the thallus ( 1.34–2.18 % N by dry wt, depending on the N-loading). In summer, the growth of both species in continuous cultures was stimulated by added ammonium. Chordaria flagelliformis showed higher rates of both growth (μmax = 0.109 · day?1) and ammonium uptake (Vmax = 23.2 μmol NH4+· h ?1 · g dry wt?1) than Fucus distichusmax = 0.074 · day?1, Vmax = 13.9 μmolNH4+ · h?1 · gdrywt?1). Calculated nitrogen subsistence quotas were similar in the two species (0.7% N in Chordaria, 0.6% N in Fucus). In those continuous cultures which did not receive added ammonium, Chordaria stored less nitrogen in excess of the subsistence quota than Fucus, although both species accumulated nitrogen at increased N-loading. Transient uptake rates were measured in response to single additions of ammonium. At any given initial concentration, short-term (30 min) uptake rates were a decreasing function of the thallus nitrogen quota and were higher for Chordaria than for Fucus. In N-depleted plants, short-term ammonium uptake rates exceeded both the uptake rates predicted from continuous cultures and the nitrogen requirements for growth. Chordaria flagelliformis, in particular, is able to scavenge efficiently low ambient N-nutrient concentrations and to sequester rapidly transient ammonium pulses. In comparison with Fucus distichus, Chordaria flagelliformis appears to be better adapted to short-term fluctuations in nutrient availability (on a scale of minutes to hours). Their N-storage capability allows both of these seaweeds to buffer the effect of fluctuations in external N-nutrient concentrations on their growth rates over periods of days to several weeks.  相似文献   

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

7.
Uptake of phosphate ions by 1 mm segments of isolated maize root cortex layers was studied. Cortex segments (from roots of 8 days old maize plants) absorb phosphate ions from 1 mM KH2PO4 in 0.2 mM CaSCO4 at the average rate of 34.3 ±3.2 μg Pi g?1 (fr. m.) h?1,i.e. 0.35± 0.02 μmol Pi g?1 (fr. m.) h?1. Phosphate uptake considerably increases after a certain period of “augmentation”,i.e. washing in aerated 0.2 mM CaSO4. This increase is completely blocked by the presence of 10 μg ml?1 cycloheximide. The relation of uptake rate to phosphate concentration in the medium was shown to have 3 phases in the concentration range of 0.02 - 40 mM. Transition points were found between 0.8–1 mM and 10–20 mM. Following Km and Vmax values were found: Km[mM] : 0.37 - 3.82 - 27.67 Vmax[μg Pi g?1 (fr. m.) h?1] : 3.33 - 39.40 - 66.67 We have found no sharp pH optimum for phosphate uptake. It proceeds at almost constant rate till pH 6.0 and then the uptake rate drops with increasing pH. At low phosphate concentrations (1 mM) the lowest uptake rate was found at 5 and 13 °C, while the uptake is higher at 5 °C than at 13 °C at phosphate concentrations higher than 1 mM. At these concentrations uptake rate at 35 °C is lower than at 25 °C. Phosphate uptake considerably decreased in anaerobic conditions. DNP and iodoacetate (0.1 mM) completely blocked phosphate uptake from 1 mM KH2PO4, while uptake from 5 and 10 mM KH2PO4 was left unaffected by these substances. The inhibitors of active - SH groups NEM and PCMB inhibited phosphate uptake: 10?3 M NEM by 81.6%, 104 M NEM by 42% and 10?4 M PCMB by 42%.  相似文献   

8.
Biodegradation of pyridine by a novel bacterial strain, Rhizobium sp. NJUST18, was studied in batch experiments over a wide concentration range (from 100 to 1,000 mg l?1). Pyridine inhibited both growth of Rhizobium sp. NJUST18 and biodegradation of pyridine. The Haldane model could be fitted to the growth kinetics data well with the kinetic constants μ* = 0.1473 h?1, K s = 793.97 mg l?1, K i = 268.60 mg l?1 and S m = 461.80 mg l?1. The true μ max, calculated from μ*, was found to be 0.0332 h?1. Yield coefficient Y X/S depended on S i and reached a maximum of 0.51 g g?1 at S i of 600 mg l?1. V max was calculated by fitting the pyridine consumption data with the Gompertz model. V max increased with initial pyridine concentration up to 14.809 mg l?1 h?1. The q S values, calculated from $V_{ \hbox{max} }$ , were fitted with the Haldane equation, yielding q Smax = 0.1212 g g?1 h?1 and q* = 0.3874 g g?1 h?1 at S m′ = 507.83 mg l?1, K s′ = 558.03 mg l?1, and K i′ = 462.15 mg l?1. Inhibition constants for growth and degradation rate value were in the same range. Compared with other pyridine degraders, μ max and S m obtained for Rhizobium sp. NJUST18 were relatively high. High K i and K i′ values and extremely high K s and K s′ values indicated that NJUST18 was able to grow on pyridine within a wide concentration range, especially at relatively high concentrations.  相似文献   

9.
The kinetics of the light-driven Cl? uptake pump of Synechococcus R-2 (PCC 7942) were investigated. The kinetics of Cl? uptake were measured in BG-11 medium (pHo, 7·5; [K+]o, 0·35 mol m?3; [Na+]o, 18 mol m?3; [Cl?]o, 0·508 mol m?3) or modified media based on the above. Net36Cl? fluxes (?Cl?o,i) followed Michaelis-Menten kinetics and were stimulated by Na+ [18 mol m?3 Na+ BG-11 ?Cl?max= 3·29±0·60 (49) nmol m?2 s?1 versus Na+-free BG-11 ?Cl?max= 1·02±0·13 (54) nmol m?2 s?1] but the Km was not significantly different in the presence or absence of Na+ at pHo 10; the Km was lower, but not affected by the presence or absence of Na+ [Km = 22·3±3·54 (20) mmol m?3]. Na+ is a non-competitive activator of net ?Cl?o,i. High [K+]o (18 mol m?3) did not stimulate net ?Cl?o,i or change the Km in Na+-free medium. High [K+]o (18 mol m?3) added to Na+ BG-11 medium decreased net ?Cl?o,i [18 mol m?3K+ BG-11; ?Cl?max= 2·50±0·32 (20) nmol m?2 s?1 versus BG-11 medium; ?Cl?max= 3·35±0·56 (20) nmol m?2 s?1] but did not affect the Km 55·8±8·100 (40) mmol m?3]. Na+-stimulation of net ?Cl?o,i followed Michaelis-Menten kinetics up to 2–5 mol m?3 [Na+]o but higher concentrations were inhibitory. The Km for Na+-stimulation of net ?Cl?o,i [K1/2(Na+)] was different at 47 mmol m?3 [Cl?]o (K1/2[Na+] = 123±27 (37) mmol m?3]. Li+ was only about one-third as effective as Na+ in stimulating Cl? uptake but the activation constant was similar [K1/2(Li+) = 88±46 (16) mmol m?3]. Br? was a competitive inhibitor of Cl? uptake. The inhibition constant (Ki) was not significantly different in the presence and absence of Na+. The overall Ki was 297±23 (45) mmol m?3. The discrimination ratio of Cl? over Br? (δCl?/δBr?) was 6·38±0·92 (df = 147). Synechococcus has a single Na+-stimulated Cl? pump because the Km of the Cl? transporter and its discrimination between Cl? and Br? are not significantly different in the presence and absence of Na+. The Cl? pump is probably driven by ATP.  相似文献   

10.
The raz1 mutant of Arabidopsis thaliana (L.) Heynh. has been selected as resistant to the toxic proline analogue, azetidine-2-carboxylic acid (2AZ). Seedlings of the mutant tolerated fivefold higher concentrations of 2AZ (ED50 = 0.25 mM) than the wild-type seedlings (ED50 = 0.05 mM). The mutant gene was found to be semi-dominant and the corresponding RAZ1 locus was mapped on chromosome 5 at 69.6±1.8 cM. The resistance to 2AZ could be fully and exclusively accounted for by the lower uptake rate of the proline analogue in the mutant. The influx of L-proline in roots of wild-type seedlings could be dissected into two components: (i) a component with a high affinity and a low capacity for l-proline (K m≈20 gmM, V max≈60 nmol·(g FW)-1·h-1) and also a high affinity for L-2AZ (K i≈40 μM) and (ii) a low-affinity, high-capacity component (K m≈5 mM: V max = 1300 nmol·(g FW)-1·h-1). Clearly, the raz1 mutation affects the activity of a high-affinity transporter, because the high-affinity uptake of proline in the mutant was at least fivefold lower than in the wild-type, whereas the low-affinity uptake was unchanged.  相似文献   

11.
Measurements of bimodal oxygen uptake have been made in a freshwater air-breathing fish,Notopterus chitala at 29.0±1(S.D.)°C. xhe mean oxygen uptake from continuously flowing water without any access to air, was found to be 3.58±0.37 (S.E.) ml O2 · h?1 and 56.84+4.29 (S.E.) ml O2 · kg?1 · h?1 for a fish weighing 66.92 + 11.27 (S.E.) g body weight. In still water with access to air, the mean oxygen uptake through the gills were recorded to be 2.49 ± 0.31 (S.E.) ml O2 · h?1 and 38.78 ± 1.92 (S.E.) ml O2 · kg?1 · h?1 and through the accessory respiratory organs (swim-bladder) 6.04±0.87 (S.E.) ml O2 · h?1 and 92.32±2.91 (S.E.) ml O2 · kg?1 · h?1 for a fish averaging 66.92±11.27 (S.E.) g. Out of the total oxygen uptake (131.10 ml O2 · kg?1 · h?1), about 70% was obtained through the aerial route and the remainder 30% through the gills.  相似文献   

12.
Porphobilinogen deaminase, the enzyme condensing four molecules of porphobilinogen, was isolated and purified from light grown Scenedesmus obliquus (wild type). The purification procedure included heat treatment, ammonium sulphate fractionation, gel filtration, high-resolution anion-exchange chromatography and hydrophobic interaction chromatography. The enzyme was purified 1368-fold, compared to the initial crude extract. Its final specific activity was 6812 units · (mg · protein)?1 at pH 7.4 with a recovery of 44%. The relative molecular mass was 33000, as determined by Sephadex G-100 gel filtration, and 35900 by lithium dodecyl sulfate-polyacrylamide-gel electrophoresis, indicating that the enzyme is a monomer. Studies of initial reaction velocities showed a linear progress curve for hydroxymethylbilane formation and a hyperbolic dependence of the initial reaction rate on substrate concentration, consistent with a sequential displacement mechanism. Apparent kinetic constants (K m and V max) for the conversion of porphobilinogen to hydroxymethylbilane at 37 ° C, pH 7.4, were 79 μM and 176 pmol · min?1, respectively. Variation of both V max and K max with pH indicated the presence of ionizable groups in the enzyme-substrate complex(es), showing a single ionization (pK 7.15) in V max/K m plots. A sharp pH-profile for V max was interpreted as a positive cooperative proton dissociation. In spite of the two pathways existing for 5-aminolevulinate biosynthesis in Scenedesmus, currently there is no indication of the existence of two porphobilinogen deaminases or even of isoenzymes.  相似文献   

13.
The response of the photosynthetic capacity (Pmax) of microphytobenthos to short-term variations of temperature (in the range 5–35° C) was assessed on a seasonal basis. The relationship is described mathematically, and relevant physiological parameters are identified: PMAX, the maximum value of Pmax achieved at Topl, the optimum temperature. Estimated values of Topt do not change significantly throughout the year and remain close to 25° C. It is thus concluded that Topt is not influenced by seasonal variations in the daily range of mud surface temperature. Identical conclusions hold for Tmax (ca. 38° C), the thermal threshold beyond which no photosynthesis occurs. Conversely, PMA estimates exhibit substantial variability: PMAX (mean ± root mean square error) is highest in April (11.18 ± 0.42 [μg C · [μg Chl a]?1· h?1) during the beginning of the annual increase in temperature, photoperiod, and maximum irradiance and is lowest in December (3.04 ± 0.16 μg C · [μg Chl a]?1· h?l). From an ecological point of view, the short-term and seasonal variations of PMAX suggest that the microphytobenthic community takes advantage of the abiotic spring environmental conditions, allowing the onset of the bloom. Nevertheless, no “acclimation strategy” (i.e. shifts in Topt and Tmax that prevent temperature inhibition in summer or improve photosynthetic rates in winter) is apparent from our results.  相似文献   

14.
Dicarboxylate transport in maize mesophyll chloroplasts   总被引:7,自引:0,他引:7  
Evidence is presented for high rates of carrier-mediated dicarboxylate anion transport in maize mesophyll chloroplasts. Radioactively labeled malate is transported across the chloroplast envelope leading to accumulation in the stroma. Malate in the stroma will exchange for external malate, oxaloacetate, glutamate, aspartate, and oxoglutarate. At 4 °C the V of malate uptake is 50 μmol·h?1·mg Chl?1 and the Km for malate is 0.5 mm. Oxaloacetate competitively inhibits malate uptake with a Ki estimated to be 0.3 mm. The temperature dependence of malate uptake indicates an activation energy of 12 kcal/mol, and extrapolation using this value gives a rate of transport at 30 °C of approximately 300 μmol·h?1·mg Chl?1. This rate approximates the rates of photosynthetic malate production by these chloroplasts.  相似文献   

15.
The effects of trace elements on ammonium degradation performance and extracellular polymeric substances (EPS) secretion of Acinetobacter harbinensis HITLi7T at low temperature were investigated. Response surface methodology (RSM) was applied to obtain the optimal composition of trace elements and analyze their correlation. In this study, the results indicated that the ammonium removal performance could be enhanced by the presence of 0.1 mg L?1 Fe, Mn, or B in pure cultivation. When the concentrations of Fe and Mn were 0.2 mg L?1, the ammonium removal rates of the novel strain HITLi7T were 0.49 ± 0.01 mg L?1·h?1 and 0.58 ± 0.01 mg L?1·h?1, respectively, while it was the low concentration of 0.05 mg L?1 B that showed the maximum ammonium removal rate (0.56 ± 0.02 mg L?1·h?1) of strain HITLi7T. The regression model was obtained and the optimal formulation of trace elements was: B 0.064 mg L?1, Fe 0.12 mg L?1, and Mn 0.1 mg L?1. Based on these values, the experimental ammonium removal rate could reach 0.59 mg L?1·h?1, which matched well with the predicted response. The study also found that the addition of trace elements, causing high ammonium removal rates, resulted in a high polysaccharide (PS) ratio in the EPS secreted by Acinetobacter harbinensis HITLi7T. Especially under the optimal conditions, the PS ratio reached the highest value of 49.9%.  相似文献   

16.
Dissolved inorganic phosphorus (DIP ) is an essential macronutrient for maintaining metabolism and growth in autotrophs. Little is known about DIP uptake kinetics and internal P‐storage capacity in seaweeds, such as Ulva lactuca (Chlorophyta). Ulva lactuca is a promising candidate for biofiltration purposes and mass commercial cultivation. We exposed U. lactuca to a wide range of DIP concentrations (1–50 μmol · L?1) and a nonlimiting concentration of dissolved inorganic nitrogen (DIN ; 5,000 μmol · L?1) under fully controlled laboratory conditions in a “pulse‐and‐chase” assay over 10 d. Uptake kinetics were standardized per surface area of U. lactuca fronds. Two phases of responses to DIP ‐pulses were measured: (i) a surge uptake (VS ) of 0.67 ± 0.10 μmol · cm?2 · d?1 and (ii) a steady state uptake (VM ) of 0.07 ± 0.03 μmol · cm?2 · d?1. Mean internal storage capacity (ISCP ) of 0.73 ± 0.13 μmol · cm?2 was calculated for DIP . DIP uptake did not affect DIN uptake. Parameters of DIN uptake were also calculated: VS  = 12.54 ± 1.90 μmol · cm?2 · d?1, VM  = 2.26 ± 0.86 μmol · cm?2 · d?1, and ISCN  = 22.90 ± 6.99 μmol · cm?2. Combining ISC and VM values of P and N, nutrient storage capacity of U. lactuca was estimated to be sufficient for ~10 d. Both P and N storage capacities were filled within 2 d when exposed to saturating nutrient concentrations, and uptake rates declined thereafter at 90% for DIP and at 80% for DIN . Our results contribute to understanding the ecological aspects of nutrient uptake kinetics in U. lactuca and quantitatively evaluating its potential for bioremediation and/or biomass production for food, feed, and energy.  相似文献   

17.
Δ53β hydroxysteroid dehydrogenase activity transforms biologically inactive Δ53β hydroxy steroids into the active Δ43-keto products (e.g. pregnenolone to progesterone). Using a cytochemical procedure which allows for the continuous microdensitometric monitoring of an enzyme reaction as it proceeds and a well described cytochemical assay for Δ53β HSD we have analysed the initial velocity rates (Vo) for dehydroepiandrosterone (DHEA) binding to this enzyme in regressing (i.e. 20α hydroxy steroid dehydrogenase positive) corpus luteum (CL) cells in unfixed tissue sections (5 μm) of the dioestrous and proestrous rat ovary. The results are mean ± S.E.M. The relationship between DHEA concentration (0 to 50 μM) and Δ53β HSD activity in the dioestrous corpora lutea was sigmoidal and had an atypical 1/Vo versus 1/S plot, the x intercept being positive. Using a 1/Vo versus 1/S2 plot the Vmax was determined to be 1·0 ± 0·08 μmol min?1 mg?1 CL (n = 6). The Hill constant was 2·7 ± 0·02 (n = 6) suggesting a high degree of positive co-operativity for DHEA binding. The S concentration for half maximal activity was 17 ± 1 μmoles (n = 6). In the corpora lutea cells of the proestrous ovary, the Vmax for DHEA transformation was unchanged (0·95 ± 0·04 μmol min?1 mg?1, n = 3) whilst the S0·5 was significantly increased to 27 ± 0·1 (p < 0·01, n = 3). The Hill constant remained positive being 2·9 ± 0·2 (n = 3). NAD+ binding to 3β HSD in regressing corpora lutea of the proestrous ovary has been demonstrated previously to be hyperbolic and fit the classical Michaelis-Menten model.1 Extending the analysis of NAD+ binding to the regressing corpus luteum of the dioestrous rat ovary revealed similar kinetic characteristics to that seen with the proestrous enzyme, the apparent Vmax and Km being 0·84 ± 0·04 μmol min?1 mg?1 CL (n = 3) and 27 ± 7 μmol 1?1 (n = 3) respectively. The Hill constant was 1·1 ± 0·03 (n = 3), indicating no co-operativity of co-factor binding.  相似文献   

18.
Ventilation frequency (FV) in motionless common sole Solea solea was measured before and after a startling stimulus in normoxia and in hypoxia (15% air saturation). Startling reduced FV in normoxia (from mean ±s.e. 41 ± 3·3 beats min?1 to near zero, i.e. 2·0 ± 1·8 beats min?1) and in hypoxia (from mean ±s.e. 80 ± 4·4 to 58·8 ± 12·9 beats min?1). It is suggested that the maintenance of high FV in hypoxia may increase the probability of detection by predators compared to normoxia.  相似文献   

19.
The kinetics of chromium(VI) reduction by Pseudomonas aeruginosa (P. aeruginosa) and Escherichia coli (E. coli) was studied under both pure and mixed cultures. Initially, the study of kinetics was performed in pure culture. It was observed that the growth of the two bacteria was both inhibited in the presence of chromium(VI). The maximum specific growth rate (μ m ) of P. aeruginosa decreased from 2.3942 h?1 (without Cr(VI)) to 1.8551 h?1 (with Cr(VI)). Under the mixed culture, the growth of E. coli was inhibited by P. aeruginosa. The maximum specific growth rate (μ m ) of E. coli decreased form 0.871 h?1 (in pure culture) to 0.153 h?1 (in mixed culture). When the concentration of each bacterium was 4.5 × 108 cells ml?1, the half-velocity reduction rate constant (K C) and the maximum specific reduction rate constant (v max) of chromium(VI) were 80.05 mg chromium(VI) l?1 and 3.674 mg chromium(VI) cells?1 h?1, respectively. The results showed that the simulation appeared in good agreement with the experimental data, supporting the series of mathematical models represented the bacteria growth and chromium(VI) reduction in both pure and mixed cultures usefully.  相似文献   

20.
Two axenic, in vitro liquid suspension cultures were established for Agardhiella subulata (C. Agardh) Kraft et Wynne, and their growth characteristics were compared. This study illustrated how reliable routes for the development of suspension cultures of macrophytic red algae of terete thallus morphology can be achieved for biotechnology applications. Undifferentiated filament clumps of 2–8 mm diameter were established by induction of callus-like tissue from thallus explants, and lightly branched microplantlets of 2–10 mm length were established by regeneration of filament clumps. The filament clumps were susceptible to regeneration. Adventitious shoot formation was reliably induced from 40% to 70% of the filament clumps by gentle mixing at 100 rev min?1 on an orbital shaker. The specific growth rate of the microplantlets was higher than the filament clumps in nonagitated well plate culture (4%–6% per day for microplantlets vs. 2%–3% per day for filament clumps) at 24° C and 8–36 μmol photons·m?2·s?1 irradiance (10:14 h LD cycle) when grown on ASP12 artificial seawater medium at pH 8.6–8.9 with 20%–25% per day medium replacement. Oxygen evolution rate vs. irradiance measurements showed that relative to the filament clumps, microplantlets had a higher maximum specific oxygen evolution rate (Po,max= 0.181 ± 0.035 vs. 0.130 ± 0.023 mmol O2·g?1 dry cell mass·h?1), but comparable respiration rate (Qo= 0.040 ± 0.013 vs. 0.033 ± 0.017 mmol O2·g?1 dry cell mass·h?1), compensation point (Ic= 3.8 ± 2.4 vs. 5.7 ± 1.2 μmol photons·m?2·s?1), and light intensity at 63.2% of saturation (Ik= 17.5 ± 3.9 vs. 14.9 ± 2.6 μmol photons·m?2·s?1). The microplantlet culture was more suitable for suspension culture development than the filament clump culture because it was morphologically stable and exhibited higher growth rates.  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