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.
Synechococcus R-2 (PCC 7942) actively accumulated Cl? in the light and dark, under control conditions (BG-11 media: pHo, 7·5; [Na+]o, 18 mol m?3; [Cl?]o, 0·508 molm?3). In BG-11 medium [Cl?], was 17·2±0·848 mol m?3 (light), electrochemical potential of Cl? (ΔμCl?i,o) =+211±2mV; [Cl?]i= 1·24±0·11 mol m?3(dark), ΔμCl?i,o=+133±4mV. Cl? fluxes, but not permeabilities, were much higher in the light: ?Cl?i,o= 4·01±5·4 nmol m?2 s?1, PCl?i,o= 47±5pm s?1 (light); ?Cl?i,o= 0·395±0·071 nmol m?2 s?1, PCl?i,o= 69±14 pm s?1 (dark). Chloride fluxes are inhibited by acid pHo (pHo 5; ?Cl?i,o= 0·14±0·04 nmol m?2 s?1); optimal at pHo 7·5 and not strongly inhibited by alkaline pHo (pHo 10; ?Cl?1i,o= 1·7±0·14 nmol m?2 s?1). A Cl?in/2H+in coporter could not account for the accumulation of Cl? alkaline pHo. Permeability of Cl? is very low, below 100pm s?1 under all conditions used, and appears to be maximal at pHo 7·5 (50–70 pm s?1) and minimal in acid pHo (20pm s?1). DCCD (dicyclohexyl-carbodiimide) inhibited ?Cl?i,o in the light about 75% and [Cl?]i fell to 2·2±0·26 (4) mol m?3. Valinomycin had no effect but monensin severely inhibited Cl? uptake ([Cl?]i= 1·02±0·32 mol m?3; ?Cl?i,o= 0·20±0·1 nmol m?2 s?1). Vanadate (200 mmol m?3) accelerated the Cl? flux (?Cl?i,o= 5·28±0·64 nmol m?2 s?1) but slightly decreased accumulation of Cl? ([Cl?], = 13·9±1·3 mol m?3) in BG-11 medium but had no significant effect in Na+-free media. DCMU (dichlorophenyldimethylurea) did not reduce [Cl?], or ?Cl?i,o to that found in the dark ([Cl?]i= 8·41±0·76 mol m?3; ?Cl?i,o= 2·06±0·36 nmol m?2 s?1). Synechococcus also actively accumulated Cl? in Na+-free media, [Cl?]i was lower but ΔΨi,o hyperpolarized in Na+-free media and so the ΔμCl?i,o was little changed ([Cl?]i= 7·98±0·698 mol m?3; ΔμCl?i,o=+203±3 mV). Net Cl? uptake was stimulated by Na+; Li+ acted as a partial analogue for Na+. Synechococcus has a Na+ activated Cl? transporter which is probably a primary 2Cl?/ATP pump. The Cl? pump is voltage sensitive. ΔμCl?i,o is directly proportional to ΔΨi,o(P»0·01%): ΔμCl?i,o= -1·487 (±0·102) ×ΔΨi,o, r= -0·983, n= 31. The ΔμCl?i,o increased (more positive) as the Δμi,o became more negative. The ΔμCl?i,o has no known function, but might provide a driving force for the uptake of micronutrients.  相似文献   

2.
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   

3.
Synechococcus R-2 is a unicellular blue-green alga. The cells will grow on Rb+ as a substitute for K+ but at a slower rate (t2~ 15 h versus 12 h). Potassium is not, strictly speaking, an essential element for Synechococcus. Rubidium duxes (using 86Rb+) are much slower than those of potassium, about 1 nmol m?2 s?1 in the light (0.35 mol m?3 Rb+). 86Rb+ fluxes in the dark are about 0.1 nmol m?2 s?1. These fluxes are very slow compared to those of Na+ and other ions. Isotopic influx of Rb+ can supply sufficient Rb+ to keep up with the demands for growth, but the net dux needed to keep up with growth in the light is a large proportion of the total observed dux. Kinetic studies of Rb+ uptake versus [Rb+] show two uptake phases consistent with a high-affinity and a low-affinity system. Both systems appear to be light-activated. Transport of Rb+ appears to be passive at pHo 10 in the light and dark. There is no case for active transport of Rb+ at pHo 7.5 in the light, but a marginal case for active uptake in the dark (about 3 kJ mol?1). There is only a small effect of Na+ upon Rb+ transport. 86Rb+ should not be used in place of 42K+ in K+ nutrition studies as the details of Rb+ transport are different to those of K+ transport.  相似文献   

4.
The semiarid and arid zones cover a quarter of the global land area and support one‐fifth of the world's human population. A significant fraction of the global soil–atmosphere exchange for climatically active gases occurs in semiarid and arid zones yet little is known about these exchanges. A study was made of the soil–atmosphere exchange of CH4, CO, N2O and NOx in the semiarid Mallee system, in north‐western Victoria, Australia, at two sites: one pristine mallee and the other cleared for approximately 65 years for farming (currently wheat). The mean (± standard error) rates of CH4 exchange were uptakes of ?3.0 ± 0.5 ng(C) m?2 s?1 for the Mallee and ?6.0 ± 0.3 ng(C) m?2 s?1 for the Wheat. Converting mallee forest to wheat crop increases CH4 uptake significantly. CH4 emissions were observed in the Mallee in summer and were hypothesized to arise from termite activity. We find no evidence that in situ growing wheat plants emit CH4, contrary to a recent report. The average CO emissions of 10.1 ± 1.8 ng(C) m?2 s?1 in the Mallee and 12.6 ± 2.0 ng(C) m?2 s?1 in the Wheat. The average N2O emissions were 0.5 ± 0.1 ng(N) m?2 s?1 from the pristine Mallee and 1.4 ± 0.3 ng(N) m?2 s?1 from the Wheat. The experimental results show that the processes controlling these exchanges are different to those in temperate systems and are poorly understood.  相似文献   

5.
Ion contents in needles from Norway spruce trees [Picea abies (L.) Karst.] growing in Würzburg and in the SO2-polluted Erzgebirge mountains were analysed to quantify cations which accumulate together with sulphate. In Würzburg there was a positive correlation of potassium (0.680 ± 0.300 Eq Eq?1 SO4?2), magnesium (0.415 ± 0.111 Eq Eq?1 SO4?2) and zinc (0.059 ± 0.006 Eq Eq?1 SO42?). In the Erzgebirge, potassium was also the stoichiometrically most important cation (0–887 ± 0–180 Eq K+ Eq?1 SO42?). All other correlations examined were weak or statistically non-significant. At both sites the calcium content of spruce needles did not depend on the sulphate content. The lack of a role for Ca2+ in neutralizing sulphate is a consequence of the presence of free oxalic acid in needles. Soluble oxalic acid precipitates Ca2+, which thereby becomes unavailable as a counterion for SO42?. The activity coefficients of Ca2+ and oxalate2?, and the solubility product of Ca-oxalate, were determined from in vivo data. It is concluded that the chronic accumulation of atmospheric sulphate in spruce needle vacuoles depletes available potassium and thereby strongly interferes with spruce growth and canopy turnover. This leads to impaired spruce vitality, even at sites where acute SO2 disease symptoms are absent.  相似文献   

6.
Shoots of poplar (Populus euramericana L. cv. Flevo) were exposed to filtered air, SO2, NH3 or a mixture of SO2 and NH3 for 7 weeks in fumigation chambers. After this exposure gas exchange measurements were carried out using a leaf chamber. As compared to leaves exposed to filtered air, leaves pretreated with 112 μg m?3 SO2 showed a small reduction in maximum CO2 assimilation rate (Pmax) and stomatal conductance (gs). They also showed a slightly higher quantum yield and dark respiration. In addition, the fluorescence measurements indicated that the Calvin cycle of the leaves pretreated with 112 μg m?3 SO2 was more rapidly activated after transition from dark to light. An exposure to 64 μg m?3 NH3 had a positive effect on Pmax, stomatal conductance and NH3 uptake of the leaves. This positive effect was counteracted by an SO2 concentration of 45 μg m?3. The exposure treatments appeared to have no effect on the relationship between net CO2-assimilation and gs. Also, no injury of the leaf cuticle or of epidermal cells was observed. Resistance analysis showed that NH3 transfer into the leaf can be estimated from data on the boundary layer and stomatal resistance for H2O transfer and NH3 concentration at the leaf surface, irrespective of whether the leaves are exposed for a short or long time to NH3 or to a mixture of NH3 and SO2. In contrast SO2 uptake into the leaves was only partly correlated to the stomatal resistance. The results suggest a large additional uptake of this gas by the leaves. The possibility of a difference in path length between SO2 and H2O molecules is proposed.  相似文献   

7.
The aim of the work was to find the optimal photon irradiance for the growth of green cells of Haematococcus pluvialis and to study the interrelations between changes in photochemical parameters and pigment composition in cells exposed to photon irradiances between 50 and 600?µmol?m?2?s?1 and a light:dark cycle of 12:12?h. Productivity of cultures increased with irradiance. However, the rate of increase was higher in the range 50–200?µmol??2?s?1. The carotenoid content increased with increasing irradiance, while the chlorophyll content decreased. The maximum quantum yield of PSII (Fv/Fm) gradually declined from 0.76 at the lowest irradiance of 50?µmol??2?s?1 to 0.66 at 600?µmol??2?s?1. Photosynthetic activity showed a drop at the end of the light period, but recovered fully during the following dark phase. A steep increase in non-photochemical quenching was observed when cultures were grown at irradiances above 200?µmol??2?s?1. A sharp increase in the content of secondary carotenoids also occurred above 200?µmol?m?2?s?1. According to our results, with H. pluvialis green cells grown in a 5-cm light path device, 200?µmol??2?s?1 was optimal for growth, and represented a threshold above which important changes in both photochemical parameters and pigment composition occurred.  相似文献   

8.
Abstract Field measurements of the gas exchange of epiphytic bromeliads were made during the dry season in Trinidad in order to compare carbon assimilation with water use in CAM and C3 photosynthesis. The expression of CAM was found to be directly influenced by habitat and microclimate. The timing of nocturnal CO2 uptake was restricted to the end of the dark period in plants found at drier habitats, and stomatal conductance in two CAM species was found to respond directly to humidity or temperature. Total night-time CO2 uptake, when compared with malic-acid formation (measured as the dawn-dusk difference in acidity, ΔH+), could only account for 10–40% of the total ΔH+ accumulated. The remaining malic acid must have been derived from the refixation of respired CO2 (recycling). Within the genus Aechmea (12 samples from four species), recycling was significantly correlated with night temperature at the six sample sites. Recycling was lowest in A. fendleri (54% of ΔH+ derived from respired CO2), a CAM bromeliad with little water-storage parenchyma that is restricted to wetter, cooler regions of Trinidad. Gas-exchange rates of C3 bromeliads were found to be similar to those of the CAM bromeliads, with CO2 uptake from 1 to 3 μmol m?2 s?1 and stomatal conductances generally up to 100 mmol m?2 s?1. The midday depression of photosynthesis occurred in exposed habitats, although photosynthetically active radiation (PAR) limited photosynthesis in shaded habitats. CO2 uptake of the C3 bromeliad Guzmania lingulata was saturated at around 500 μmol m?2 s?1 PAR, suggesting that epiphytic plants found in the shaded forest understorey are shade-tolerant rather than shade-demanding. Transpiration ratios (TR) during CO2 fixation in CAM (Phase I and IV) and C3 bromeliads were compared at different sites in order to assess the efficiency of water utilization. For the epiphytes displaying marked uptake of CO2, TR were found to be lower than many previously published values. In addition, the average TR values were very similar for dark CO2 uptake in CAM (42 ± 41, n= 12), Phase IV of CAM (69 ± 36, n= 3) and for C3 photosynthesis (99 ± 73, n= 4) in these plants. It appears that recycling of respired CO2 by CAM bromeliads and efficient use of water in all phases of CO2 uptake are physiological adaptations of bromeliads to arid microclimates in the humid tropics.  相似文献   

9.
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   

10.
The dry matter production in Polytrichum commune protonemata was increased when the light intensity was increased from 0 to 160 μE m?2 s?1, and at 160 μE m?2 s?1 production was about 200% of that found at 17 μE m?2 s?1. Production of chlorophyll (Chl) was increased by increasing light intensity from 0 to 17 μE m?2 s?1, but decreasing at light intensities above 17 μE m?2 s?1. At 160 μE m?2 s?1 the production of Chl was only about 50% of that at 17 μE m?2 s?1. The rate of CO2 fixation was low (0.31 μg CO2/mg Chi × h) at the light intensity of 17 μE m?2 s?1 as compared with that at 160 μE m?2 s?1 (0.83 μg CO2/mg Chi × h). Production of mono- (MGDG) and diglycosyl diglycerides (DGDG) was closely associated with that of chlorophylls. At the higher light intensity (160 μE m?2 s?1) production of glycolipids was about 60% of that at 17 μE m?2 s?1. Production of more polar lipids was less affected by light intensity. Light intensity also affected the fatty acid pattern of the lipid fractions. The effect was most pronounced in the MGDG fraction, where the proportion of C 18: 3ω3 + C 16: 3ω3 was higher at the higher light intensity.  相似文献   

11.
The effects of salinity, light intensity and sediment on Gracilaria tenuistipitata C.F. Chang & B.M. Xia on growth, pigments, agar production, and net photosynthesis rate were examined in the laboratory under varying conditions of salinity (0, 25 and 33 psu), light intensity (150, 400, 700 and 1000 µmol photons m?2 s?1) and sediment (0, 0.67 and 2.28 mg L?1). These conditions simulated field conditions, to gain some understanding of the best conditions for cultivation of G. tenuistipitata. The highest growth rate was at 25 psu, 700 µmol photons m?2 s?1 with no sediments, that provided a 6.7% increase in weight gain. The highest agar production (24.8 ± 3.0 %DW) was at 25 psu, 150–400 µmol photons m?2 s?1 and no sediment. The highest pigment contents were phycoerythrin (0.8 ± 0.5 mg g?1FW) and phycocyanin (0.34 ± 0.05 mg g?1 FW) produced in low light conditions, at 150 µmol photons m?2 s?1. The highest photosynthesis rate was 161.3 ± 32.7 mg O2 g?1 DW h?1 in 25 psu, 400 µmol photons m?2 s?1 without sediment in the short period of cultivation, (3 days) and 60.3 ± 6.7 mg O2 g?1 DW h?1 in 25 psu, 700 µmol photons m?2 s?1 without sediment in the long period of cultivation (20 days). The results indicated that salinity was the most crucial factor affecting G. tenuistipitata growth and production. This would help to promote the cultivation of Gracilaria cultivation back into the lagoon using these now determined baseline conditions. Extrapolation of the results from the laboratory study to field conditions indicated that it was possible to obtain two crops of Gracilaria a year in the lagoon, with good yields of agar, from mid‐January to the end of April (dry season), and from mid‐July to the end of September (first rainy season) when provided sediment was restricted.  相似文献   

12.
Switchgrass (Panicum virgatum L.) has gained importance as feedstock for bioenergy over the last decades due to its high productivity for up to 20 years, low input requirements, and potential for carbon sequestration. However, data on the dynamics of CO2 exchange of mature switchgrass stands (>5 years) are limited. The objective of this study was to determine net ecosystem exchange (NEE), ecosystem respiration (Re), and gross primary production (GPP) for a commercially managed switchgrass field in its sixth (2012) and seventh (2013) year in southern Ontario, Canada, using the eddy covariance method. Average NEE flux over two growing seasons (emergence to harvest) was ?10.4 μmol m?2 s?1 and reached a maximum uptake of ?42.4 μmol m?2 s?1. Total annual NEE was ?380 ± 25 and ?430 ± 30 g C m?2 in 2012 and 2013, respectively. GPP reached ?1354 ± 23 g C m?2 in 2012 and ?1430 ± 50g C m?2 in 2013. Annual Re in 2012 was 974 ± 20 g C m?2 and 1000 ± 35 g C m?2 in 2013. GPP during the dry year of 2012 was significantly lower than that during the normal year of 2013, but yield was significantly higher in 2012 with 1090 g  m?2, compared to 790 g m?2 in 2013. If considering the carbon removed at harvest, the net ecosystem carbon balance came to 106 ± 45 g C  m?2 in 2012, indicating a source of carbon, and to ?59 ± 45 g C m?2 in 2013, indicating a sink of carbon. Our results confirm that switchgrass can switch between being a sink and a source of carbon on an annual basis. More studies are needed which investigate this interannual variability of the carbon budget of mature switchgrass stands.  相似文献   

13.
In isolated Elodea densa leaves, the relationships between H+ extrusion (-ΔH+), K+ fluxes and membrane potential (Em) were investigated for two different conditions of activation of the ATP-dependent H+ pump. The ‘basal condition’ (darkness, no pump activator present) was characterized by low values of-ΔH+ and K+ uptake (ΔK+), wide variability of the ?ΔH+/ΔK+ ratio, relatively low membrane polarization and Em values more positive than EK for external K+ concentrations (|K+]o of up to 2mol m?3. A net K+ uptake was seen already at [K+]o below 1 mol m?3, suggesting that K+ influx in this condition was a thermodynamically uphill process involving an active mechanism. When the H+ pump was stimulated by fusicoccin (FC), by cytosol acidification, or by light (the ‘high polarization condition’), K+ influx largely dominated K+ and C? efflux, and the ?ΔH+/ΔK+ ratio approached unity. In the range 50 mmol m?3?5 mol m?3 [K+]0, Em was consistently more negative than EK. The curve of K+ influx at [K+]0 ranging from 50 to 5000mmol m?3 fitted a monophasic, hyperbolic curve, with an apparent half saturation value = 0–2 mol m?3. Increasing |K+]0 progressively depolarized Em, counteracting the strong hyperpolarizing effect of FC. The effects of K+ in depolarizing Em were well correlated with the effects on both K+ influx and ?ΔH+, suggesting a cause-effect chain: K+0 influx → depolarization → activation of H+ extrusion. Cs+ competitively inhibited K+ influx much more strongly in the ‘high polarization’ than in the ‘basal’ condition (50% inhibition at [Cs+]/[K+]0 ratios of 1:14 and 1:2, respectively) thus confirming the involvement of different K+ uptake systems in the two conditions. These results suggest that in E. densa leaves two distinct modes of interactions rule the relationships between H+ pump, membrane polarization and K+ transport. At low membrane polarization, corresponding to a low state of activation of the PM H+-ATPase and to Em values more positive than EK, K+ influx would mainly  相似文献   

14.
The aim of this study was to investigate the effect of time-of-day on Preferred Transition Speed (PTS) and spatiotemporal organization of walking and running movements. Twelve active male subjects participated in the study (age: 27.2?±?4.9 years; height: 177.9?±?5.4?cm; body mass: 75.9?±?5.86?kg). First, PTS was determined at 08:00?h and 18:00?h. The mean of the two PTS recorded at the two times-of-day tested was used as a reference (PTSm). Then, subjects were asked to walk and run on a treadmill at three imposed speeds (PTSm, PTSm?+?0.3?m.s?1, and PTSm???0.3?m.s?1) at 08:00?h and 18:00?h. Mean stride length, temporal stride, spatial stride variability, and temporal stride variability were used for gait analysis. The PTS observed at 08:00?h (2.10?±?0.17?m.s?1) tends to be lower (p?=?0.077) than that recorded at 18:00?h (2.14?±?0.19?m.s?1). Stride lengths recorded while walking (p?=?0.038) and running (p?=?0.041) were shorter at 08:00?h than 18:00?h. No time-of-day effect was observed for stride frequency during walking and running trials. When walking, spatial stride variability (p?=?0.020) and temporal stride variability (p?=?0.028) were lower at 08:00?h than at 18:00?h. When running, no diurnal variation of spatial stride variability or temporal stride variability was detected.  相似文献   

15.
Smooth muscle cells isolated from the secondary and tertiary branches of the rabbit mesenteric artery contain large Ca2+-dependent channels. In excised patches with symmetrical (140 mm) K+ solutions, these channels had an average slope conductance of 235 ± 3 pS, and reversed in direction at −6.1 ± 0.4 mV. The channel showed K+ selectivity and its open probability (P o ) was voltage-dependent. Iberiotoxin (50 nm) reversibly decreased P o , whereas tetraethylammonium (TEA, at 1 mm) reduced the unitary current amplitude. Apamin (200 nm) had no effect. The channel displayed sublevels around 1/3 and 1/2 of the mainstate level. The effect of [Ca2+] on P o was studied and data fitted to Boltzmann relationships. In 0.1, 0.3, 1.0 and 10 μm Ca2+, V 1/2 was 77.1 ± 5.3 (n= 18), 71.2 ± 4.8 (n= 16), 47.3 ± 10.1 (n= 11) and −14.9 ± 10.1 mV (n= 6), respectively. Values of k obtained in 1 and 10 μm [Ca2+] were significantly larger than that observed in 0.1 μm [Ca2+]. With 30 μm NS 1619 (a BKCa channel activator), V 1/2 values were shifted by 39 mV to the left (hyperpolarizing direction) and k values were not affected. TEA applied intracellularly, reduced the unitary current amplitude with a K d of 59 mm. In summary, BKCa channels show a particularly weak sensitivity to intracellular TEA and they also display large variation in V 1/2 and k. These findings suggest the possibility that different types (isoforms) of BKCa channels may exist in this vascular tissue. Received: 22 December 1997/Revised: 27 March 1998  相似文献   

16.
When expressed in Xenopus oocytes KAAT1 increases tenfold the transport of l-leucine. Substitution of NaCl with 100 mm LiCl, RbCl or KCl allows a reduced but significant activation of l-leucine uptakes. Chloride-dependence is not strict since other pseudohalide anions such as thyocyanate are accepted. KAAT1 is highly sensitive to pH. It can transport l-leucine at pH 5.5 and 8, but the maximum uptake has been observed at pH 10, near to the physiological pH value, when amino and carboxylic groups are both deprotonated. The pH value mainly influences the V max in Na+ activation curves and l-leucine kinetics. The kinetic parameters are K mNa = 4.6 ± 2 mm, V maxNa = 14.8 ± 1.7 pmol/oocyte/5 min for pH 8.0 and K mNa = 2.8 ± 0.7 mm, V maxNa = 31.3 ± 1.9 pmol/oocyte/5 min for pH 10.0. The kinetic parameters of l-leucine uptake are: K m = 120.4 ± 24.2 μm, V max = 23.2 ± 1.4 pmol/oocyte/5 min at pH 8.0 and K m = 81.3 ± 24.2 μm, V max = 65.6 ± 3.9 pmol/oocyte/5 min at pH 10.0. On the basis of inhibition experiments, the structural features required for KAAT1 substrates are: (i) a carboxylic group, (ii) an unsubstituted α-amino group, (iii) the side chain is unnecessary, if present it should be uncharged regardless of length and ramification. Received: 27 April 1999/Revised: 10 January 2000  相似文献   

17.
Abstract: There is debate about the mechanisms mediating adenosine release from neurons. In this study, the release of adenosine evoked by depolarizing cultured cerebellar granule neurons with 50 mM K+ was inhibited by 49 ± 7% in Ca2+-free medium. The remaining release was blocked by dipyridamole (IC50 = 6.4 × 10?8M) and nitrobenzylthioinosine (IC50 = 3.6 × 10?8M), inhibitors of adenosine uptake. Ca2+-dependent release was reduced by 78 ± 9% following a 21-h pretreatment of the cells with pertussis toxin, which ADP-ribosylates Gi/Go G proteins, thereby preventing their dissociation. The nucleoside transporter-mediated component of K+-induced adenosine release also was inhibited by 62 ± 8% by pertussis toxin and was potentiated by 78 ± 11% following cholera toxin treatment, which permanently activates Gs. Uptake of [3H]adenosine into cultured cerebellar granule neurons over a 10-min period was not dependent on extracellular Na+ but was reduced by dipyridamole (IC50 = 3.2 × 10?8M) and nitrobenzylthioinosine (IC50 = 2.6 × 10?8M). Thus, adenosine uptake likely occurs via the same transporter mediating Ca2+-independent adenosine release. Adenosine uptake was potentiated by cholera toxin pretreatment (152 ± 15% of control), but pertussis toxin had no statistically significant effect. It is possible that Gs, Gi/Go, or free Gβγ dimer modulate the equilibrative, inhibitor-sensitive nucleoside carrier to enhance adenosine transport.  相似文献   

18.
Abstract Some characteristics of photosynthetic inorganic carbon uptake by Palmaria palmata, a marine red macroalga, have been measured under physiological conditions in artificial seawater. The apparent affinity of thallus for CO2 [K1/2(CO2)] at pH 8.0 and 15°C was 21.4±3.0mmol m?3 CO2 under air, and 25.7±70mmol m?3 CO2 under N2. The corresponding values of Vmax were 2.98 ± 0.42 and 3.65±0.87 mmol O2 evolved g Chr?1 s?l. The apparent Km(CO2) of isolated ribulose bisphosphate carboxylase was determined at pH 8.0 and 30 °C to be 30.2 mmol m?3 CO2, and the corresponding value of Vmax was 19.67 μniol CO2 g protein?1 s?1. The CO2 compensation points of the thallus were measured in artificial seawater at pH 8.0 under air and N2, using a gas-chromatographic method. The values were relatively low, rising from 10 cm3 m?3 at 15°C, to 35 cm3 m?3 at 25°C, but were not affected by the O2 concentration. The lack of an effect of O2 on photosynthesis and on compensation point indicates that there is little photorespiratory CO2 loss in this macroalga. The high affinity of the thallus for CO2, and the low CO2 compensation concentrations, are consistent with the occurrence of bicarbonate uptake in this alga.  相似文献   

19.
Kim K  Lee X 《Plant, cell & environment》2011,34(10):1790-1801
Dew formation, a common meteorological phenomenon, is expected to intensify in the future. Dew can influence the H218O and HDO isotopic compositions of leaf water (δL), but the phenomenon has been neglected in many experimental and modelling studies. In this study, the dew effect on δL was investigated with a dark plant chamber in which dew formation was introduced. The H218O and HDO compositions of water vapour, dew water and leaf water of five species were measured for up to 48 h of dew exposure. Our results show that the exchanges of H218O and HDO in leaf water with the air continued in the darkness when the net H216O flux was zero. Our estimates of the leaf conductance using the isotopic mass balance method ranged from 0.035 to 0.087 mol m?2 s?1, in broad agreement of the night‐time stomatal conductance reported in the literature. In our experiments, the conductance of the C4 species was 0.04 ± 0.01 mol m?2 s?1 and that of the C3 plants was 0.10 ± 0.04 mol m?2 s?1. At the end of 16 h dew exposure, 72 (±17) and 94 (±11)% of the leaf water came from dew according to the 18O and D tracer, respectively.  相似文献   

20.
The effect of CO2 on potassium transport by Chlorella fusca   总被引:1,自引:1,他引:0  
Abstract. The effect of CO2 on net K+ uptake by Chlorella fusca grown on high CO2 levels was examined by passing 1.5% CO2 through algal suspensions gassed previously with air or CO2-free air Addition of CO2 in the light caused a large net uptake of K+ (initial velocity 4.2–9.2 mmol s?1 m?3 cells) which decreased the concentration of K+ in the supernatant from 0.1–0.2 mol m?3 to 3–10 mmol m?3. In the dark and in the presence of 30 mmol m?3 DCMU, no effects were found. Measurement or the unidirectional K+ fluxes by using 86Rb+ as a label showed that in the presence of 1.5% CO2, influx of K+ was increased by a factor of 2–4 while efflux was inhibited completely. CO2 hyperpolarized the membrane potential (determined through TPP+ uptake) from –120mV to –130 mV which could not explain the more than 15,000-fold K+ accumulations. In the light, CO2 lowered the intracellular pH (determined with DMO) by 0.5 units. In the dark and in the presence of DCMU only, a small acidification of 0.1 units was found. During the first 15 min after addition of CO2 the malate content of the cells increased from 0.7 to 1.5 mol m?3 packed cells. On the basis of these and earlier results, CO2-induced net K+ uptake is interpreted as a stimulation of an electroneutral ATP-dependent K+/H+ exchange at the plasmalemma. This exchange acts as a ‘pHstat’ by reducing the intracellular acidification caused by production of acidic assimilation products.  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