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 140 毫秒
1.
Summary Chlamydomonas reinhardtii cells provide an effective system for glycolate photoproduction, operative during 30 h when they are growing under low CO2, in the presence of 1 mM aminooxyacetate and 50 M acetazolamide. Glycolate excretion by the cells can proceed for about 4 days if every other 12 h a high CO2 level is restored in the culture in the absence of inhibitors. The immobilized system in alginate beads has about a twofold higher glycolate photoproduction rate (23 mol·mg chlorophyll (Chl)–1·h–1) than free-living cells (12 mol · mg Chl–1 · h–1). Offprint requests to: C. Vílchez  相似文献   

2.
Nitrate and nitrite was reduced by Escherichia coli E4 in a l-lactate (5 mM) limited culture in a chemostat operated at dissolved oxygen concentrations corresponding to 90–100% air saturation. Nitrate reductase and nitrite reductase activity was regulated by the growth rate, and oxygen and nitrate concentrations. At a low growth rate (0.11 h–1) nitrate and nitrite reductase activities of 200 nmol · mg–1 protein · min–1 and 250 nmol · mg–1 protein · min–1 were measured, respectively. At a high growth rate (0.55 h–1) both enzyme activities were considerably lower (25 and 12 nmol mg–1 · protein · min–1). The steady state nitrite concentration in the chemostat was controlled by the combined action of the nitrate and nitrite reductase. Both nitrate and nitrite reductase activity were inversely proportional to the growth rate. The nitrite reductase activity decreased faster with growth rate than the nitrate reductase. The chemostat biomass concentration of E. coli E4, with ammonium either solely or combined with nitrate as a source of nitrogen, remained constant throughout all growth rates and was not affected by nitrite concentrations. Contrary to batch, E. coli E4 was able to grow in continuous cultures on nitrate as the sole source of nitrogen. When cultivated with nitrate as the sole source of nitrogen the chemostat biomass concentration is related to the activity of nitrate and nitrite reductase and hence, inversely proportional to growth rate.  相似文献   

3.
The presence and properties of the enzymes involved in the synthesis and conversion of phospho(enol)pyruvate (PEP) and oxaloacetate (OAA), the precursors for aspartate-derived amino acids, were investigated in three different Corynebacterium strains. This study revealed the presence of both PEP carboxykinase 0.29 mol·min–1·mg–1 of protein [units (U)·mg–1] and PEP synthetase (0.13 U·mg–1) in C. 2 glutamicum as well as pyruvate kinase (1.4 U·mg–1) and PEP carboxylase (0.16 U·mg–1). With the exception of PEP carboxykinase these activities were also present in glucose-grown C. flavum and C. lactofermentum. Pyruvate carboxylase activity was not detected in all three species cultivated on glucose or lactate. At least five enzyme activities that utilize OAA as a substrate were detected in crude extracts of C. glutamicum: citrate synthase (2 U·mg–1), malate dehydrogenase (2.5 U·mg–1), glutamate: OAA transaminase (1 U·mg–1), OAA-decarboxylating activity (0.89 U·mg–1) and the previously mentioned PEP carboxykinase (0.29 U·mg–1). The partially purified OAA-decarboxylase activity of C. glutamicum was completely dependent on the presence of inosine diphosphate and Mn2+, had a Michaelis constant (K m) of 2.0mm for OAA and was inhibited by ADP and coenzyme A (CoA). Examination of the kinetic properties showed that adenine nucleotides and CoA derivatives have reciprocal but reinforcing effects on the enzymes catalyzing the interconversion of pyruvate, PEP and OAA in C. glutamicum. A model for the regulation of the carbon flow based on these findings is presented.Correspondence to: M. S. M. Jetten  相似文献   

4.
We have compared the characteristics of nitrate uptake by Aphanothece halophytica grown under non-stress and salt-stress conditions. Both cell types showed essentially similar patterns of nitrate uptake toward ammonium, nitrite, and DL-glyceraldehyde. Although the affinities of nitrate to non-stress cells and salt-stress cells were not significantly different, i.e., Ks = 416 and 450 µM, respectively, the Vmax value for non-stress cells was about twofold of that for salt-stress cells (9.1 vs 5.3 µmol min–1 mg–1 Chl). Nitrate uptake by A. halophytica was found to be dependent on Na+. Ammonium inhibited nitrate uptake, and the presence of methionine sulfoximine could not release the inhibition by ammonium. Nitrite appeared to competitively inhibit nitrate uptake with a Ki value of 84 µM. Both chloride and phosphate anions did not affect nitrate uptake. DL-Glyceraldehyde, an inhibitor of CO2 fixation, caused a reduction in the uptake of nitrate.Received: 22 October 2002 / Accepted: 6 December 2002  相似文献   

5.
Summary Na+–H+ exchange activity in renal brush border membrane vesicles isolated from hyperthyroid rats was increased. When examined as a function of [Na+], treatment altered the initial rate of Na+ uptake by increasingV m (hyperthyroid, 18.9±1.1 nmol Na+ · mg–1 · 2 sec–1; normal, 8.9±0.3 nmol Na+ · mg–1 · 2 sec–1), and not the apparent affinityK Na + (hyperthyroid, 7.3±1.7mm; normal, 6.5±0.9mm). When examined as a function of [H+] and at a subsaturating [Na+] (1mm), hyperthyroidism resulted in the proportional increase in Na+ uptake at every intravesicular pH measured. A positive cooperative effect on Na+ uptake was found with increased intravesicular acidity in vesicles from both normal and hyperthyroid rats. When the data were analyzed by the Hill equation, it was found that hyperthyroidism did not change then (hyperthyroid, 1.2±0.06; normal, 1.2±0.07) or the [H+]0.5 (hyperthyroid, 0.39±0.08 m; normal, 0.44±0.07 m) but increased the apparentV m (hyperthyroid, 1.68±0.14 nmol Na+ · mg–1 · 2 sec–1; normal 0.96±0.10 nmol Na+ · mg–1 · 2 sec–1). The uptake of Na+ in exchange for H+ in membrane vesicles from normal and hyperthyroid animals was not influenced by membrane potential. H+ translocation or debinding was rate limiting for Na+–H+ exchange since Na+–Na+ exchange activity was greater than Na+–H+ exchange activity. Hyperthyroidism caused a proportional increase and hypothyroidism caused a proportional decrease in Na+–Na+ and Na+–H+ exchange. We conclude that hyperthyroidism leads to either an increase in the number of functional exchangers in the membrane or exactly proportional increases in the rate-limiting steps for Na+–Na+ and Na+–H+ exchange activity.  相似文献   

6.
Two Leuconostoc oenos mutant strains unable to metabolize malic acid were differentiated by [U-14C]-labelled L-malate transport assays into a malolactic-enzyme-deficient mutant and a malate-transport-defective mutant. A mathematical analysis of the data from L-malic acid uptake at three pH values (5.2, 4.5, and 3.2) in the malolactic-enzyme-deficient strains suggest two simultaneous uptake mechanisms, presumably a carrier-mediated transport and a passive diffusion for the anionic and the undissociated forms of the acid, respectively. The apparent affinity constant (K m t) and the maximal rate (V m t) values for L-malate active transport were, 12 mM and 43 mol L-malate·mg–1·s–1, respectively. Active transport was constitutive and strongly inhibited by protonophores and by ATPase inhibitors. L-Lactic acid appeared to inhibit L-malic acid transport, suggesting an L-lactate/L-malate exchange. At pH values of 4.5 or above, the passive diffusion of L-malic acid was negligible. However, at pH 3.2, the mean pH of wine, the permeability of the cells to the undissociated acid by simple diffusion could represent more than 50% of total L-malic acid uptake, with a diffusion constant (K D) of 0.1 s–1. Correspondence to: C. Divies  相似文献   

7.
Summary Free-living or immobilized Chlamydomonas reinhardtii cells photoproduce ammonium from nitrite in a medium containing 1 mM of l-methionine-d,l-sulphoximine (MSX). Ammonium is accumulated in the medium to 8 mM final concentration, which inhibits nitrite uptake by the MSX-treated cells and consequently the excretion of ammonium is blocked. However, if ammonium was removed from the medium and nitrite and MSX periodically restored, the photoproduction process could be maintained over 96 h, with a final ammonium concentration of about 18 mM for free-living cells and 28 mM for immobilized ones. The MSX-treated cells showed a photoproduction productivity of 1300 mol NH 4 + · mg chlorophyll (Chl)-1, with an average production rate of 14 mol NH 4 + · mg Chl-1 per hour, for calcium alginate-entrapped cells, while the corresponding data for free-living ones was 650 mol NH 4 + · mg Chl-1 and 6.7 mol NH 4 + · mg Chl-1 per hour, respectively. Immobilized cells showed a significant increase in the nitrite uptake rate, probably due to a change in membrane permeability as a consequence of cell-matrix interactions.  相似文献   

8.
Summary Measurements of unidirectional calcium fluxes in stripped intestinal epithelium of the tilapia,Oreochromis mossambicus, in the presence of ouabain or in the absence of sodium indicated that calcium absorption via the fish intestine is sodium dependent. Active Ca2+ transport mechanisms in the enterocyte plasma membrane were analyzed. The maximum capacity of the ATP-dependent Ca2+ pump (V m :0.63 nmol·min–1 mg–1,K m : 27nm Ca2+) is calculated to be 2.17 nmol·min–1·mg–1, correcting for 29% inside-out oriented vesicles in the membrane preparation. The maximum capacity of the Na+/Ca2+ exchanger with high affinity for Ca2+ (V m :7.2 nmol·min–1·mg–1,K m : 181nm Ca2+) is calculated to be 13.6 nmol·min–1·mg–1, correcting for 53% resealed vesicles and assuming symmetrical behavior of the Na+/Ca2+ exchanger. The high affinity for Ca2+ and the sixfold higher capacity of the exchanger compared to the ATPase suggest strongly that the Na+/Ca2+ exchanger will contribute substantially to Ca2+ extrusion in the fish enterocyte. Further evidence for an important contribution of Na+/Ca2+ exchange to Ca2+ extrusion was obtained from studies in which the simultaneous operation of ATP-and Na+-gradient-driven Ca2+ pumps in inside-out vesicles was evaluated. The fish enterocyte appears to present a model for a Ca2+ transporting cell, in which Na+/Ca2+ exchange activity with high affinity for Ca2+ extrudes Ca2+ from the cell.  相似文献   

9.
Summary Chlamydomonas reinhardtii cells immobilized in Ba-alginate beads provide a stable and effective system for photoproducing ammonium from nitrite in a culture medium containing l-methionine-d,l-sulphoximine. The process was studied in either air-lift, fluidized or packed-bed reactors, the last one providing the most suitable system with a volumetric activity of 2700 mol NH inf+ sup4 ·1–1 per hour.  相似文献   

10.
The CO2 production of individual larvae of Apis mellifera carnica, which were incubated within their cells at a natural air humidity of 60–80%, was determined by an open-flow gas analyzer in relation to larval age and ambient temperature. In larvae incubated at 34 °C the amount of CO2 produced appeared to fall only moderately from 3.89±1.57 µl mg–1 h–1 in 0.5-day-old larvae to 2.98±0.57 µl mg–1 h–1 in 3.5-day-old larvae. The decline was steeper up to an age of 5.5 days (0.95±1.15 µl mg–1 h–1). Our measurements show that the respiration and energy turnover of larvae younger than about 80 h is considerably lower (up to 35%) than expected from extrapolations of data determined in older larvae. The temperature dependency of CO2 production was determined in 3.5-day-old larvae, which were incubated at temperatures varying from 18 to 38 °C in steps of 4 °C. The larvae generated 0.48±0.03 µl mg–1 h–1 CO2 at 18 °C, and 3.97±0.50 µl mg–1 h–1 CO2 at 38 °C. The temperature-dependent respiration rate was fitted to a logistic curve. We found that the inflection point of this curve (32.5 °C) is below the normal brood nest temperature (33–36 °C). The average Q10 was 3.13, which is higher than in freshly emerged resting honeybees but similar to adult bees. This strong temperature dependency enables the bees to speed up brood development by achieving high temperatures. On the other hand, the results suggest that the strong temperature dependency forces the bees to maintain thermal homeostasis of the brood nest to avoid delayed brood development during periods of low temperature.Abbreviations m body mass - R rate of development or respiration - TI inflexion point of a logistic (sigmoid) curve - TL lethal temperature - TO temperature of optimum (maximum) developmentCommunicated by G. Heldmaier  相似文献   

11.
Summary A double-chambered bioreactor based on a composite immobilized-cell gel layer/microporous membrane structure was applied to the continuous denitrification of high-nitrate water. Immobilized denitrifying bacteria (Pseudomonas denitrificans) were provided with separate flows of nitrate and carbon (C) nutrient, with no contamination of the treated water by cell leakage from the gel. Using acetate (7.5 mm) as a C source and a C/N ratio of 3 (mol/mol), specific denitrification rates ranging from 15 to 25 g NO inf3 sup– · h–1 · – cm–2 membrane surface (50–85 g NO inf3 sup– · h–1 · cm–3 gel) were obtained. The denitrifying activity remained stable for several months. At the flow rate used (10 cm3 · h–1), the effluents contained noticeable amounts of NO inf2 sup– ions but the treated water remained uncontaminated by the carbon nutrient. Most NO inf2 sup– ions disappeared from the treated water in a second reactor connected in series. When fed with an unchlorinated sludge supernatant as C nutrient, immobilized bacteria performed efficient denitrification of water for only 3 weeks. Diffusion experiments showed that acetate ions diffused much less rapidly than NO inf3 sup– or NO inf2 sup– ions through the composite structure. Further developments of the system are considered.  相似文献   

12.
Cell-free extracts of crotonate-grown cells of the syntrophic butyrate-oxidizing bacteriumSyntrophospora bryantii contained high hydrogenase activities (8.5–75.8 µmol · min–1 mg–1 protein) and relatively low formate dehydrogenase activities (0.04–0.07 µmol · min–1 mg–1 protein). The K M value and threshold value of the hydrogenase for H2 were 0.21 mM and 18 µM, respectively, whereas the K M value and threshold value of the formate dehydrogenase for formate were 0.22 mM and 10 µM, respectively. Hydrogenase, butyryl-CoA dehydrogenase and 3-OH-butyryl-CoA dehydrogenase were detected in the cytoplasmic fraction. Formate dehydrogenase and CO2 reductase were membrane-bound, likely located at the outer aspect of the cytoplasmic membrane. Results suggest that during syntrophic butyrate oxidation H2 is formed intracellularly while formate is formed at the outside of the cell.  相似文献   

13.
Summary Geotrichum candidum (isolate 1–9) pathogenic on citrus fruits, appears to lack siderophore production. Iron uptake byG. candidum is mediated by two distinct iron-regulated, energy-and temperature-dependent transport systems that require sulfhydryl groups. One system exhibits specificity for either ferric or ferrous iron, whereas the other exhibits specificity for ferrioxamine-B-mediated iron uptake and presumably other hydroxamate siderophores. Radioactive iron uptake from59FeCl3 showed an optimum at pH 6 and 35° C, and Michaelis-Menten kinetics (apparentK m = 3 m,V max = 0.054 nmol · mg–1 · min–1). The maximal rate of Fe2+ uptake was higher than Fe3+ (V max = 0.25 nmol · mg–1 · min–1) but theK m was identical. Reduction of ferric to ferrous iron prior to transport could not be detected. The ferrioxamine B system exhibits an optimum at pH 6 and 40° C and saturation kinetics (K m = 2 M,V max = 0.22 nmol · mg–1 · min–1). The two systems were distinguished as two separate entities by negative reciprocal competition, and on the basis of differential response to temperature and phenazine methosulfate. Mössbauer studies revealed that cells fed with either57FeCl3 or57FeCl2 accumulated unknown ferric and ferrous binding metabolites.  相似文献   

14.
The conversion of glycerol to 1,3-propanediol by Citrobacter freundii DSM 30040 was optimized in single- and two-stage continuous cultures. The productivity of 1,3-propanediol formation was highest under glycerol limitation and increased with the dilution rate (D) to a maximum of 3.7 g·l–1·h–1. Glycerol dehydratase seemed to be the rate-limiting step in 1,3-propanediol formation. Conditions for the two-stage fermentation process were as follows: first stage, glycerol limitation (250mM), pH 7.2, D=0.1 h, 31° C; second stage, additional glycerol, pH 6.6, D=0.05 h–1, 28° C. Under these conditions 875mM glycerol were consumed, the final 1,3-propanediol concentration was 545mM, and the overall productivity 1.38 g·1–1·h–1. Correspondence to: G. Gottschalk  相似文献   

15.
Newrkla  P.  Gunatilaka  A. 《Hydrobiologia》1982,91(1):531-536
Benthic community respiration rates of profundal sediments of Fuschlsee (37.6 mg · O2 · m–2 · h–1 — eutrophic), Mondsee (40.19 mg · O2 · m–2 · h–1 — eutrophic) and Attersee (11.5 mg · O2 · m–2 · h–1 — oligo-mesotrophic) were measuredin situ, and in cores. By exposing the sediments to different oxygen levels in the laboratory it was found that benthic community metabolism reduced with decreasing oxygen concentrations. The slope of the regression lines, relating oxygen uptake rates to oxygen concentrations, differed significantly for the different sites investigated. These results were closely related to the trophic conditions of the lakes.  相似文献   

16.
The concentration dependence of the influx ofl-lysine in excised roots ofArabidopsis thaliana seedlings was analyzed for the wild-type (WT) and two mutants,rlt11 andraec1, which had been selected as resistant to lysine plus threonine, and to S-2-aminoethyl-l-cysteine, respectively. In the WT three components were resolved: (i) a high-affinity, low-capacity component [K m = 2.2 M;V max = 23 nmol·(g FW)–1·h–1]; (ii) a low-affinity, high-capacity component [K m = 159 M;V max = 742 nmol·(g FW)–1·h–1]; (iii) a component which is proportional to the external concentration, with a constant of proportionalityk = 104 nmol·(g FW)–1 h–1];·mM–1. The influx ofl-lysine in the mutants was lower than in the WT, notably in the concentration range 0.1–0.4 mM, where it was only 7% of that in the WT. In both mutants the reduced influx could be fully attributed to the absence of the low-affinity (high-K m ) component. This component most likely represents the activity of a specific basic-amino-acid transporter, since it was inhibited by several other basic amino acids (arginine, ornithine, hydroxylysine, aminoethylcysteine) but not byl-valine. The high-affinity uptake ofl-lysine may be due to the activity of at least two general amino acid transporters, as it was inhibitable byl-valine, and could be further dissected into two components with a high affinity (K i = 1–5 M; and a low affinity (K i = 0.5–1mM) forl-valine, respectively. Therlt11 andraecl mutant have the same phenotype and the corresponding loci were mapped on chromosome 1, but it is not yet clear whether they are allelic.Abbreviations AEC S-2-aminoethyl-l-cysteine - K i equilibrium constant - WT wild-type  相似文献   

17.
Nitrogen fixation was measured in four subarctic streams substantially modified by beaver (Castor canadensis) in Quebec. Acetylene-ethylene (C2H2 C2H4) reduction techniques were used during the 1982 ice-free period (May–October) to estimate nitrogen fixation by microorganisms colonizing wood and sediment. Mean seasonal fixation rates were low and patchy, ranging from zero to 2.3 × 10–3 µmol C2H4 · cm–2 · h–1 for wood, and from zero to 7.0 × 10–3 µmol C2H4 · g AFDM–1 · h–1 for sediment; 77% of all wood and 63% of all sediment measurements showed no C2H2 reduction. Nonparametric statistical tests were unable to show a significant difference (p > 0.05) in C2H2 reduction rates between or within sites for wood species or by sediment depth.Nitrogen contributed by microorganisms colonizing wood in riffles of beaver influenced watersheds was small (e.g., 0.207 g N · m–2 · y–1) but greater than that for wood in beaver ponds (e.g., 0.008 g N · m–2 · y–1) or for streams without beaver (e.g., 0.003 g N · m–2 · y–1). Although mass specific nitrogen fixation rates did not change significantly as beaver transform riffles into ponds, the nitrogen fixed by organisms colonizing sediment in pond areas (e.g., 5.1 g N · m–2 · y–1) was greater than that in riffles (e.g., 0.42 g N · m–2 · y–1). The annual nitrogen contribution is proportional to the amount of sediment available for microbial colonization. We estimate that total nitrogen accumulation in sediment, per unit area, is enhanced 9 to 44 fold by beaver damming a section of stream.  相似文献   

18.
The uptake of ammonium, nitrate and phosphate by laboratory-grown young sporophytes of Laminaria abyssalis was measured in a perturbed system (batch mode) at 18 °C and 35 ± 5 µE m–2 s–1 photon flux density. Uptake of all appeared to follow saturation-type nutrient uptake kinetics. The NO inf3 sup– (K s = 14.0 µM, V max = 5.0 µmol h–1 g–1 dry wt) and NH inf4 sup+ (K s = 4.6 µM, V max= 2.0 µmol h–1 g–1 dry wt) were taken up simultaneously, although NH inf4 sup+ was taken up more rapidly. Values of K 3 and V max for phosphate were, respectively, 2.21 µM and 0.83 µmol h–1 g–1 dry wt. Nitrate and phosphate were both consumed in similar rates (V max /Ks 0.37) at low concentrations. NH inf4 sup+ , thus, might be a more efficient form of N fertilizer if artificial enrichment of seawater is used.  相似文献   

19.
Desulfotomaculum acetoxidans has been proposed to oxidize acetate to CO2 via an oxidative acetyl-CoA/carbon monoxide dehydrogenase pathway rather than via the citric acid cycle. We report here the presence of the enzyme activities required for the operation of the novel pathway. In cell extracts the following activities were found (values in brackets=specific activities and apparent K m; 1 U·mg-1=1 mol·min-1·mg protein-1 at 37°C): Acetate kinase (6.3 U·mg-1; 2 mM acetate; 2.4 mM ATP); phosphate acetyltransferase (60 U·mg-1, 0.4 mM acetylphosphate; 0.1 mM CoA); carbon monoxide dehydrogenase (29 U·mg-1; 13% carbon monoxide; 1.3 mM methyl viologen); 5,10-methylenetetrahydrofolate reductase (3 U·mg-1, 0.06 mM CH3–FH4); methylenetetrahydrofolate dehydrogenase (3.6 U·mg-1, 0.9 mM NAD, 0.1 mM CH2=FH4); methenyltetrahydrofolate cyclohydrolase (0.3 U·mg-1); formyltetrahydrofolate synthetase (3 U·mg-1, 1.4 mM FH4, 0.4 mM ATP, 13 mM formate); and formate dehydrogenase (10 U·mg-1, 0.4 mM formate, 0.5 mM NAD). The specific activities are sufficient to account for the in vivo acetate oxidation rate of 0.26 U·mg-1.Non-standard abbreviations FH4 Tetrahydrofolate - CHO-FH4 N10-formyltetrahydrofolate - CHFH4 N5,N10-methenyltetrahydrofolate - CH2=FH4 N5,N10-methylenetetrahydrofolate - CH3–FH4 N5-methyltetrahydrofolate - MOPS morpholinopropane sulfonic acid - DTT d,l-1,4-dithiothreitol - TRIS tris-(hydroxymethyl)-aminomethane - Ap5A p1,P5-di(adenosine-5)pentaphosphate - MV methyl viologen  相似文献   

20.
Summary Pulmonary CO-diffusing capacity (D l CO), lung volume, pulmonary perfusion and O2-uptake were measured by non-invasive techniques in the lizardsVaranus exanthematicus andTupinambis teguixin (mean body weight 2.2 kg for both species).The CO-diffusing capacity was at 25–27°C 0.059 mlstpd·kg–1·min–1·Torr–1 inVaranus, which is 47% greater than the value of 0.040 mlstpd·kg–1·min–1·Torr–1 inTupinambis. The lung volume ofVaranus was 36 ml·kg–1 and that ofTupinambis 20 ml·kg–1. At 35–37°C the diffusing capacity of lizard lungs are about 25% of those for mammals of comparable size.InVaranus pulmonary CO-diffusing capacity increased with temperature from 0.027 mlstpd·kg–1·min–1·Torr–1 at 17–19 °C to 0.075 mlstpd·kg–1·min–1·Torr–1 at 35–37 °C. This change closely matched a concomitant increase of O2-uptake. Pulmonary perfusion increased from 27 ml·kg–1·min–1 to 55 ml·kg–1·min–1 within this temperature range.The study emphasizes that pulmonary diffusing capacity cannot be fully evaluated without information on pulmonary perfusion and O2-uptake. In reptiles and other ectotherms diffusing capacity must be reported at specified body temperature.  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