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.
Nutrient uptake by forest trees is dependent on ectomycorrhizal (EM) mycelia that grow out into the soil from the mycorrhizal root tips. We estimated the production of EM mycelia in root free samples of pure spruce and mixed spruce-oak stands in southern Sweden as mycelia grown into sand-filled mesh bags placed at three different soil depths (0–10, 10–20 and 20–30 cm). The mesh bags were collected after 12 months and we found that 590±70 kg ha–1 year–1 of pure mycelia was produced in spruce stands and 420±160 kg ha–1 year–1 in mixed stands. The production of EM mycelia in the mesh bags decreased with soil depth in both stand types but tended to be more concentrated in the top soil in the mixed stands compared to the spruce stands. The fungal biomass was also determined in soil samples taken from different depths by using phospholipid fatty acids as markers for fungal biomass. Subsamples were incubated at 20°C for 5 months and the amount of fungal biomass that degraded during the incubation period was used as an estimate of EM fungal biomass. The EM biomass in the soil profile decreased with soil depth and did not differ significantly between the two stand types. The total EM biomass in the pure spruce stands was estimated to be 4.8±0.9×103 kg ha–1 and in the mixed stands 5.8±1.1×103 kg ha–1 down to 70 cm depth. The biomass and production estimates of EM mycelia suggest a very long turnover time or that necromass has been included in the biomass estimates. The amount of N present in EM mycelia was estimated to be 121 kg N ha–1 in spruce stands and 187 kg N ha–1 in mixed stands. The 13C value for mycelia in mesh bags was not influenced by soil depth, indicating that the fungi obtained all their carbon from the tree roots. The 13C values in mycelia collected from mixed stands were intermediate to values from pure spruce and pure oak stands suggesting that the EM mycelia received carbon from both spruce and oak trees in the mixed stands. The 15N value for the EM mycelia and the surrounding soil increased with soil depth suggesting that they obtained their entire N from the surrounding soil.  相似文献   

2.
Balnokin YV  Popova LG  Pagis LY  Andreev IM 《Planta》2004,219(2):332-337
Our previous investigations have established that Na+ translocation across the Tetraselmis viridis plasma membrane (PM) mediated by the primary ATP-driven Na+-pump, Na+-ATPase, is accompanied by H+ counter-transport [Y.V. Balnokin et al. (1999) FEBS Lett 462:402–406]. The hypothesis that the Na+-ATPase of T. viridis operates as an Na+/H+ exchanger is tested in the present work. The study of Na+ and H+ transport in PM vesicles isolated from T. viridis demonstrated that the membrane-permeant anion NO3 caused (i) an increase in ATP-driven Na+ uptake by the vesicles, (ii) an increase in (Na++ATP)-dependent vesicle lumen alkalization resulting from H+ efflux out of the vesicles and (iii) dissipation of electrical potential, , generated across the vesicle membrane by the Na+-ATPase. The (Na++ATP)-dependent lumen alkalization was not significantly affected by valinomycin, addition of which in the presence of K+ abolished at the vesicle membrane. The fact that the Na+-ATPase-mediated alkalization of the vesicle lumen is sustained in the absence of the transmembrane is consistent with a primary role of the Na+-ATPase in driving H+ outside the vesicles. The findings allowed us to conclude that the Na+-ATPase of T. viridis directly performs an exchange of Na+ for H+. Since the Na+-ATPase generates electric potential across the vesicle membrane, the transport stoichiometry is mNa+/nH+, where m>n.Abbreviations BTP Bis-Tris-Propane, 1,3-bis[tris(hydroxymethyl)methylamino]-propane - CCCP Carbonyl cyanide m-chlorophenylhydrazone - DTT Dithiothreitol - NCDC 2-Nitro-4-carboxyphenyl N,N-diphenylcarbamate - PMSF Phenylmethylsulfonyl fluoride - PM Plasma membrane  相似文献   

3.
Gallbladder Na+ absorption is linked to gallstone formation in prairie dogs. We previously reported Na+/H+ exchanger (NHE1-3) expression in native gallbladder tissues. Here we report the functional characterization of NHE1, NHE2 and NHE3 in primary cultures of prairie dog gallbladder epithelial cells (GBECs). Immunohistochemical studies showed that GBECs grown to confluency are homogeneous epithelial cells of gastrointestinal origin. Electron microscopic analysis of GBECs demonstrated that the cells form polarized monolayers characterized by tight junctions and apical microvilli. GBECs grown on Snapwells exhibited polarity and developed transepithelial short-circuit current, Isc, (11.6 ± 0.5 µA · cm–2), potential differences, Vt (2.1 ± 0.2 mV), and resistance, Rt (169 ± 12 · cm2). NHE activity in GBECs assessed by measuring dimethylamiloride-inhibitable 22Na+ uptake under a H+ gradient was the same whether grown on permeable Snapwells or plastic wells. The basal rate of 22Na+ uptake was 21.4 ± 1.3 nmol · mg prot–1 · min–1, of which 9.5 ± 0.7 (~45%) was mediated through apically-restricted NHE. Selective inhibition with HOE-694 revealed that NHE1, NHE2 and NHE3 accounted for ~6%, ~66% and ~28% of GBECs total NHE activity, respectively. GBECs exhibited saturable NHE kinetics (Vmax 9.2 ± 0.3 nmol · mg prot–1 · min–1; Km 11.4 ± 1.4 mM Na+). Expression of NHE1, NHE2 and NHE3 mRNAs was confirmed by RT-PCR analysis. These results demonstrate that the primary cultures of GBECs exhibit Na+ transport characteristics similar to native gallbladder tissues, suggesting that these cells can be used as a tool for studying the mechanisms of gallbladder ion transport both under physiologic conditions and during gallstone formation.  相似文献   

4.
The lobster (Homarus americanus) hepato-pancreatic epithelial baso-lateral cell membrane possesses three transport proteins that transfer calcium between the cytoplasm and hemolymph: an ATP-dependent calcium ATPase, a sodium-calcium exchanger, and a verapamil-sensitive cation channel. We used standard centrifugation methods to prepare purified hepato-pancreatic baso-lateral membrane vesicles and a rapid filtration procedure to investigate whether 65Zn2+ transfer across this epithelial cell border occurs by any of these previously described transporters for calcium. Baso-lateral membrane vesicles were osmotically reactive and exhibited a time course of uptake that was linear for 10–15 s and approached equilibrium by 120 s. In the absence of sodium, 65Zn2+ influx was a hyperbolic function of external zinc concentration and followed the Michaelis-Menten equation for carrier transport. This carrier transport was stimulated by the addition of 150 M ATP (increase in Km and Jmax) and inhibited by the simultaneous presence of 150 mol l–1 ATP+250 mol l–1 vanadate (decrease in both Km and Jmax). In the absence of ATP, 65Zn2+ influx was a sigmoidal function of preloaded vesicular sodium concentration (0, 5, 10, 20, 30, 45, and 75 mmol l–1) and exhibited a Hill Coefficient of 4.03±1.14, consistent with the exchange of 3 Na+/1Zn2+. Using Dixon analysis, calcium was shown to be a competitive inhibitor of baso-lateral membrane vesicle 65Zn2+ influx by both the ATP-dependent (Ki=205 nmol l–1 Ca2+) and sodium-dependent (Ki=2.47 mol l–1 Ca2+) transport processes. These results suggest that zinc transport across the lobster hepato-pancreatic baso-lateral membrane largely occurred by the ATP-dependent calcium ATPase and sodium-calcium exchanger carrier proteins.Communicated by: I.D. Hume  相似文献   

5.
Superfusion of heart cells with hyperosmotic solution causes cell shrinkage and inhibition of membrane ionic currents, including delayed-rectifer K+ currents. To determine whether osmotic shrinkage also inhibits inwardly-rectifying K+ current (IK1), guinea-pig ventricular myocytes in the perforated-patch or ruptured-patch configuration were superfused with a Tyrodes solution whose osmolarity (T) relative to isosmotic (1T) solution was increased to 1.3–2.2T by addition of sucrose. Hyperosmotic superfusate caused a rapid shrinkage that was accompanied by a negative shift in the reversal potential of Ba2+-sensitive IK1, an increase in the amplitude of outward IK1, and a steepening of the slope of the inward IK1-voltage (V) relation. The magnitude of these effects increased with external osmolarity. To evaluate the underlying changes in chord conductance (GK1) and rectification, GK1-V data were fitted with Boltzmann functions to determine maximal GK1 (GK1max) and voltage at one-half GK1max (V0.5). Superfusion with hyperosmotic sucrose solutions led to significant increases in GK1max (e.g., 28±2% with 1.8T), and significant negative shifts in V0.5 (e.g., –6.7±0.6 mV with 1.8T). Data from myocytes investigated under hyperosmotic conditions that do not induce shrinkage indicate that GK1max and V0.5 were insensitive to hyperosmotic stress per se but sensitive to elevation of intracellular K+. We conclude that the effects of hyperosmotic sucrose solutions on IK1 are related to shrinkage-induced concentrating of intracellular K+.  相似文献   

6.
The Ca2+-conducting pathway of myocytes isolated from the cricket lateral oviduct was investigated by means of the whole-cell patch clamp technique. In voltage-clamp configuration, two types of whole cell inward currents were identified. One was voltage-dependent, initially activated at –40 mV and reaching a maximum at 10 mV with the use of 140 mM Cs2+-aspartate in the patch pipette and normal saline in the bath solution. Replacement of the external Ca2+ with Ba2+ slowed the current decay. Increasing the external Ca2+ or Ba2+ concentration increased the amplitude of the inward current and the current–voltage (I–V) relationship was shifted as expected from a screening effect on negative surface charges. The inward current could be carried by Na+ in the absence of extracellular Ca2+. Current carried by Na+ (I Na) was almost completely blocked by the dihydropyridine Ca2+ channel antagonist, nifedipine, suggesting that the I Na is through voltage-dependent L-type Ca2+ channels. The other inward current is voltage-independent and its I–V relationship was linear between –100 mV to 0 mV with a slight inward rectification at more hyperpolarizing membrane potentials when 140 mM Cs+-aspartate and 140 mM Na+-gluconate were used in the patch pipette and in the bath solution, respectively. A similar current was observed even when the external Na+ was replaced with an equimolar amount of K+ or Cs+, or 50 mM Ca2+ or Ba2+. When the osmolarity of the bath solution was reduced by removing mannitol from the bath solution, the inward current became larger at negative potentials. The I–V relationship for the current evoked by the hypotonic solution also showed a linear relationship between –100 mV to 0 mV. Bath application of Gd3+ (10 M) decreased the inward current activated by membrane hyperpolarization. These results clearly indicate that the majority of current activated by a membrane hyperpolarization is through a stretch-activated Ca2+-permeable nonselective cation channel (NSCC). Here, for the first time, we have identified voltage-dependent L-type Ca2+ channel and stretch-activated Ca2+-permeable NSCCs from enzymatically isolated muscle cells of the cricket using the whole-cell patch clamp recording technique.Abbreviations I Ca Ca2+ current - I Na Na+ current - I–V current–voltage - NSCC nonselective cation channel Communicated by G. Heldmaier  相似文献   

7.
From Bacillus subtilis cell extracts, ferredoxin-NADP+ reductase (FNR) was purified to homogeneity and found to be the yumC gene product by N-terminal amino acid sequencing. YumC is a 94-kDa homodimeric protein with one molecule of non-covalently bound FAD per subunit. In a diaphorase assay with 2,6-dichlorophenol-indophenol as electron acceptor, the affinity for NADPH was much higher than that for NADH, with Km values of 0.57 M vs >200 M. Kcat values of YumC with NADPH were 22.7 s–1 and 35.4 s–1 in diaphorase and in a ferredoxin-dependent NADPH-cytochrome c reduction assay, respectively. The cell extracts contained another diaphorase-active enzyme, the yfkO gene product, but its affinity for ferredoxin was very low. The deduced YumC amino acid sequence has high identity to that of the recently identified Chlorobium tepidum FNR. A genomic database search indicated that there are more than 20 genes encoding proteins that share a high level of amino acid sequence identity with YumC and which have been annotated variously as NADH oxidase, thioredoxin reductase, thioredoxin reductase-like protein, etc. These genes are found notably in gram-positive bacteria, except Clostridia, and less frequently in archaea and proteobacteria. We propose that YumC and C. tepidum FNR constitute a new group of FNR that should be added to the already established plant-type, bacteria-type, and mitochondria-type FNR groups.  相似文献   

8.
Measurement of nitrogen isotopic composition (15N) of plants and soil nitrogen might allow the characteristics of N transformation in an ecosystem to be detected. We tested the measurement of 15N for its ability to provide a picture of N dynamics at the ecosystem level by doing a simple comparison of 15N between soil N pools and plants, and by using an existing model. 15N of plants and soil N was measured together with foliar nitrate reductase activity (NRA) and the foliar NO3 pool at two sites with different nitrification rates in a temperature forest in Japan. 15N of plants was similar to that of soil NO3 in the high-nitrification site. Because of high foliar NRA and the large foliar NO3 pool at this site, we concluded that plant 15N indicated a great reliance of plants on soil NO3 there. However, many 15N of soil N overlapped each other at the other site, and 15N could not provide definitive evidence of the N source. The existing model was verified by measured 15N of soil inorganic N and it explained the variations of plant 15N between the two sites in the context of relative importance of nitrification, but more information about isotopic fractionations during plant N uptake is required for quantitative discussions about the plant N source. The model applied here can provide a basis to compare 15N signatures from different ecosystems and to understand N dynamics.  相似文献   

9.
We used the patch-clamp technique to identify and characterize the electrophysiological, biophysical, and pharmacological properties of K+ channels in enzymatically dissociated ventricular cells of the land pulmonate snail Helix. The family of outward K+ currents started to activate at –30 mV and the activation was faster at more depolarized potentials (time constants: at 0 mV 17.4 ± 1.2 ms vs. 2.5 ± 0.1 ms at + 60 mV). The current waveforms were similar to those of the A-type family of voltage-dependent K+ currents encoded by Kv4.2 in mammals. Inactivation of the current was relatively fast, i.e., 50.2 ± 1.8% of current was inactivated within 250 ms at + 40 mV. The recovery of K+ channels from inactivation was relatively slow with a mean time constant of 1.7 ± 0.2 s. Closer examination of steady-state inactivation kinetics revealed that the voltage dependency of inactivation was U-shaped, exhibiting less inactivation at more depolarized membrane potentials. On the basis of this phenomenon, we suggest that a channel encoded by Kv2.1 similar to that in mammals does exist in land pulmonates of the Helix genus. Outward currents were sensitive to 4-aminopyridine and tetraethylammonium chloride. The last compound was most effective, with an IC50 of 336 ± 142 µmol l–1. Thus, using distinct pharmacological and biophysical tools we identified different types of voltage-gated K+ channels. Present address for S.A.K.: Brigham and Womens Hospital, Cardiovascular Division, Harvard Medical School, 75 Francis St., Thorn 1216, Boston, MA 02115.  相似文献   

10.
Production of hydrogen peroxide has been found in Ulva rigida (Chlorophyta). The formation of H2O2 was light dependent with a production of 1.2 mol·g FW–1·h–1 in sea water (pH 8.2) at an irradiance of 700 mol photons m–2·s–1. The excretion was also pH dependent: in pH 6.5 the production was not detectable (< 5 nmol·g FW–1·h–1) but at pH 9.0 the production was 5.0 mol·g FW–1·h–1. The production of H2O2 was totally inhibited by 3-(3,4-dichlorophenyl)-1,1 dimethylurea (DCMU). The ability of U. rigida growing in tanks (7501) under a natural light regime to excrete H2O2 was checked and found to be seven times higher at 08.00 hours than other times of the day. The H2O2 concentration in the cultivation tank (density: 2 g FW·l–1) reached the highest value (3 M) at 11.00 hours. Photosynthesis was not influenced by H2O2 formation. The H2O2 is suggested to come from the Mehler reaction (pseudocyclic photophosphorylation). With an oxygen evolution of 120 mmol·g FW–1·h–1 at pH 8.2 and 90 mmol·g FW–1·h–1 at pH 9.0, 0.5% and 2.7% of the electrons were used for extracellular H2O2 production. The H2O2 production is sufficiently high to be of physiological and ecological significance, and is suggested to be a part of the defence against epi and endophytes.Abbreviations ACL artificial, continuous light - DCMU 3-(3,4-dichlorophenyl)-1,1-dimethylurea - GNL greenhouse - LDC Luminol-dependent chemiluminescence - SOD Superoxide dismutase This investigation was supported by SAREC (Swedish Agency for Research Cooperation with Developing Countries), Hierta-Retzius Foundation, Marianne and Marcus Wallenberg Foundation, the Swedish Environmental Protection Board, and CICYT Spain.  相似文献   

11.
This study investigates the reverse mode of the Na+/glucose cotransporter (SGLT1). In giant excised inside-out membrane patches from Xenopus laevis oocytes expressing rabbit SGLT1, application of α-methyl-D-glucopyranoside (αMDG) to the cytoplasmic solution induced an outward current from cytosolic to external membrane surface. The outward current was Na+- and sugar-dependent, and was blocked by phlorizin, a specific inhibitor of SGLT1. The current-voltage relationship saturated at positive membrane voltages (30–50 mV), and approached zero at −150 mV. The half-maximal concentration for αMDG-evoked outward current (K0.5αMDG) was 35 mM (at 0 mV). In comparison, K0.5αMDG for forward sugar transport was 0.15 mM (at 0 mV). K0.5Na was similar for forward and reverse transport (≈35 mM at 0 mV). Specificity of SGLT1 for reverse transport was: αMDG (1.0) > D-galactose (0.84) > 3-O-methyl-glucose (0.55) > D-glucose (0.38), whereas for forward transport, specificity was: αMDG ≈ D-glucose ≈ D-galactose > 3-O-methyl-glucose. Thus there is an asymmetry in sugar kinetics and specificity between forward and reverse modes. Computer simulations showed that a 6-state kinetic model for SGLT1 can account for Na+/sugar cotransport and its voltage dependence in both the forward and reverse modes at saturating sodium concentrations. Our data indicate that under physiological conditions, the transporter is poised to accumulate sugar efficiently in the enterocyte.  相似文献   

12.
This study addresses the mechanisms of oxygen-induced regulation of ion transport pathways in mouse erythrocyte, specifically focusing on the role of cellular redox state and ATP levels. Mouse erythrocytes possess Na+/K+ pump, K+-Cl and Na+-K+-2Cl cotransporters that have been shown to be potential targets of oxygen. The activity of neither cotransporter changed in response to hypoxia-reoxygenation. In contrast, the Na+/K+ pump responded to hypoxic treatment with reversible inhibition. Hypoxia-induced inhibition was abolished in Na+-loaded cells, revealing no effect of O2 on the maximal operation rate of the pump. Notably, the inhibitory effect of hypoxia was not followed by changes in cellular ATP levels. Hypoxic exposure did, however, lead to a rapid increase in cellular glutathione (GSH) levels. Decreasing GSH to normoxic levels under hypoxic conditions abolished hypoxia-induced inhibition of the pump. Furthermore, GSH added to the incubation medium was able to mimic hypoxia-induced inhibition. Taken together these data suggest a pivotal role of intracellular GSH in oxygen-induced modulation of the Na+/K+ pump activity.  相似文献   

13.
The isolation of a soluble brain fraction which behaves as an endogenous ouabain-like substance, termed endobain E, has been described. Endobain E contains two Na+, K+-ATPase inhibitors, one of them identical to ascorbic acid. Neurotransmitter release in the presence of endobain E and ascorbic acid was studied in non-depolarizing (0 mM KCl) and depolarizing (40 mM KCl) conditions. Synaptosomes were isolated from cerebral cortex of male Wistar rats by differential centrifugation and Percoll gradient. Synaptosomes were preincubated in HEPES-saline buffer with 1 mM d-[3H]aspartate (15 min at 37°C), centrifuged, washed, incubated in the presence of additions (60 s at 37°C) and spun down; radioactivity in the supernatants was quantified. In the presence of 0.5–5.0 mM ascorbic acid, d-[3H]aspartate release was roughly 135–215% or 110–150%, with or without 40 mM KCl, respectively. The endogenous Na+, K+-ATPase inhibitor endobain E dose-dependently increased neurotransmitter release, with values even higher in the presence of KCl, reaching 11-times control values. In the absence of KCl, addition of 0.5–10.0 mM commercial ouabain enhanced roughly 100% d-[3H]aspartate release; with 40 mM KCl a trend to increase was recorded with the lowest ouabain concentrations to achieve statistically significant difference vs. KCl above 4 mM ouabain. Experiments were performed in the presence of glutamate receptor antagonists. It was observed that MPEP (selective for mGluR5 subtype), failed to decrease endobain E response but reduced 50–60% ouabain effect; LY-367385 (selective for mGluR1 subtype) and dizocilpine (for ionotropic NMDA glutamate receptor) did not reduce endobain E or ouabain effects. These findings lead to suggest that endobain E effect on release is independent of metabotropic or ionotropic glutamate receptors, whereas that of ouabain involves mGluR5 but not mGluR1 receptor subtype. Assays performed at different temperatures indicated that in endobain E effect both exocytosis and transporter reversion are involved. It is concluded that endobain E and ascorbic acid, one of its components, due to their ability to inhibit Na+, K+-ATPase, may well modulate neurotransmitter release at synapses.  相似文献   

14.
This study examined the osmoregulatory status of the euryhaline elasmobranch Carcharhinus leucas acclimated to freshwater (FW) and seawater (SW). Juvenile C. leucas captured in FW (3 mOsm l–1 kg–1) were acclimated to SW (980–1,000 mOsm l–1 kg–1) over 16 days. A FW group was maintained in captivity over a similar time period. In FW, bull sharks were hyper-osmotic regulators, having a plasma osmolarity of 595 mOsm l–1 kg–1. In SW, bull sharks had significantly higher plasma osmolarities (940 mOsm l–1 kg–1) than FW-acclimated animals and were slightly hypo-osmotic to the environment. Plasma Na+, Cl, K+, Mg2+, Ca2+, urea and trimethylamine oxide (TMAO) concentrations were all significantly higher in bull sharks acclimated to SW, with urea and TMAO showing the greatest increase. Gill, rectal gland, kidney and intestinal tissue were taken from animals acclimated to FW and SW and analysed for maximal Na+/K+-ATPase activity. Na+/K+-ATPase activity in the gills and intestine was less than 1 mmol Pi mg–1 protein h–1 and there was no difference in activity between FW- and SW-acclimated animals. In contrast Na+/K+-ATPase activity in the rectal gland and kidney were significantly higher than gill and intestine and showed significant differences between the FW- and SW-acclimated groups. In FW and SW, rectal gland Na+/K+-ATPase activity was 5.6±0.8 and 9.2±0.6 mmol Pi mg–1 protein h–1, respectively. Na+/K+-ATPase activity in the kidney of FW and SW acclimated animals was 8.4±1.1 and 3.3±1.1 Pi mg–1 protein h–1, respectively. Thus juvenile bull sharks have the osmoregulatory plasticity to acclimate to SW; their preference for the upper reaches of rivers where salinity is low is therefore likely to be for predator avoidance and/or increased food abundance rather than because of a physiological constraint.  相似文献   

15.
13C-13C NOESY experiments were performed under long mixing time conditions on reduced human superoxide dismutase (32 kDa, 15N, 13C and 70% 2H labeled). 13C-13C couplings were successfully eliminated through post-processing of in-phase-anti-phase (IPAP) data. It appears that at mixing time m of 3.0 s the spin diffusion mechanism allows the detection of 96% of the two-bond correlations involving C and C. The interpretation was confirmed by simulations. This approach broadens the range of applicability of 13C-13C NOESY spectroscopy.  相似文献   

16.
A. Laisk  O. Kiirats  V. Oja  U. Gerst  E. Weis  U. Heber 《Planta》1992,186(3):434-441
Exchange of CO2 and O2 and chlorophyll fluorescence were measured in the presence of 360 1 · 1–1 CO2 in nitrogen in Helianthus annuss L. leaves which had been preconditioned in the dark or at a photon flux density (PFD) of 24 mol · m–2 · s–1 either in 21 or 0% O2. An initial light-dependent O2 outburst of 6 mol · m–2 was measured after aerobic dark incubation. It was attributed to the reduction of electron carriers, predominantly plastoquinone. The maximum initial rate of O2 evolution at PFD 8000 mol · m–2 · s–1 was 170 mol · m–2 · s–2 or about four times the steady CO2-and light-saturated rate of photosynthesis. Fluorescence measurements showed that the rate was still acceptor-limited. Fast O2 evolution ceased after electron carriers were reduced in the dark-adapted leaf, but continued for a short time at the lower rate of 62 mol · m–2 · s–1 in the light-adapted leaf. The data are interpreted to show that enzymes involved in 3-phosphoglycerate reduction are dark-inhibited, but were fully active in low light. In a dark-adapted leaf, respiratory CO2 evolution continued under nitrogen; it was partially inhibited by illumination. Prolonged exposure of a leaf to anaerobic conditions caused reducing equivalents to accumulate. This was shown by a slowly increasing chlorophyll fluorescence yield which indicated the reduction of the PSII acceptor QA in the dark. When the leaf was illuminated, no O2 evolution was detected from short light pulses, although transient O2 production was appreciable during longer light pulses. This indicates that an electron donor (pool size about 2–3 e/PSII reaction center) became reduced in the dark and the first photons were used to oxidise this donor instead of water.Abbreviations Chl chlorophyll - CRC carbon reduction cycle - GAPDH NADP-glyceraldehyde-phosphate dehydrogenase - PFD photon flux density - PGA 3-phosphoglycerate - RuBP ribulose bisphosphate - TCA tricarboxylic acid cycle To whom correspondence should be addressedThis work received support by the Estonian Academy of Sciences, the Gottfried-Wilhelm-Leibniz Program of the Deutsche For-schungsgemeinschaft and the Sonderforschungsbereich 251 of the University of Würzburg.  相似文献   

17.
Two major indicators were used to access the degree of photorespiration in various photosynthetic types of Flaveria species (C3, C3-C4, C4-like, and C4): the O2 inhibition of photosynthesis measured above the O2 partial pressure which gives a maximum rate, and O2- and light-dependent whole-chain electron flow measured at the CO2 compensation point (). The optimum level of O2 for maximum photosynthetic rates under atmospheric levels of CO2 (34 Pa) was lower in C3 and C3-C4 species (ca. 2 kPa) than in C4-like and C4 species (ca. 9 kPa). Increasing O2 partial pressures from the optimum for photosynthesis up to normal atmospheric levels (ca. 20 kPa) caused an inhibition of photosynthesis which was more severe under lower CO2. This inhibition was calculated as the O2 inhibition index (A, the percentage inhibition of photosynthesis per kPa increase in O2). From measurements of 18 Flaveria species at atmospheric CO2, the A values decreased from C3 (1.9–2.1) to C3-C4 (1.2–1.6), C4-like (0.6–0.8) and C4 species (0.3–0.4), indicating a progressive decrease in apparent photorespiration in this series. With increasing irradiance at under atmospheric levels of O2, and increasing O2 partial pressure at 300 mol quanta·m–2·s–1, there was a similar increase in the rate of O2 evolution associated with whole-chain electron flow (Jo 2, calculated from chlorophyll fluorescence analysis) in the C3 and C3-C4 species compared to a much lower rate in the C4-like and C4 species. The results indicate that there is substantial O2-dependent electron flow in C3 and C3-C4 species, reflecting a high level of photorespiration compared to that in C4-like and C4 species. Consistent with these results, there was a significant decrease in from C3 (6–6.2 Pa) to C3-C4 (1.0–3.0 Pa), to C4-like and C4 species (0.3–0.8 Pa), indicating a progressive decrease in apparent photorespiration. However, C3 and C3-C4 species examined had high intrinsic levels of photorespiration with the latter maintaining low apparent rates of photorespiration and lower values, primarily by refixing photorespired CO2. The C4-like and C4 Flaveria species had low, but measurable, levels of photorespiration via selective localization of ribulose-1,5-bisphosphate carboxylase in bundle sheath cells and operation of a CO2 pump via the C4 pathway.Abbreviations and Symbols A CO2 assimilation rate - CE carboxylation efficiency - Ci intercellular CO2 partial pressure - Ia absorbed PPFD - Jo 2 oxygen evolution from PSII - PPFD photosynthetic photon flux density (mol · m–2· s–1) - Rubisco ribulose-1,5-bisphosphate carboxylase/oxygenase - RuBP ribulose-1,5-bisphosphate - VPD water-vapor pressure difference between the leaf and atmospheric air - CO2 compensation point - CO 2 quantum yield of CO2 assimilation - PSII quantum yield of photosystem II - A O2 inhibition index for photosynthesis (percentage inhibition of photosynthesis per kPa increase in O2) This research was supported by the National Science Foundation Grant IBN 9317756 and Equipment (Grant DMB-8515521 and DOE/USDA/NSF Triagency Plnat Biochemistry Research Training Grant Program.  相似文献   

18.
Degradation of membrane phospholipids is associated with apoptotic responses, but the signaling development of this degradation is not well understood. Cerium (Ce4+), an important rare earth element, induces cellular apoptosis and taxol biosynthesis in Taxus cuspidata suspension cultures. Here, using mass spectrometry and biochemical technique, we demonstrated that the phospholipase D (PLD) was rapidly activated by Ce4+ and hydrolyzed structural phospholipids to generate lipid signal molecule, phosphatidic acid (PA). 1-Butanol, an antagonist of PLD-dependent PA production, blocked the biphasic burst of superoxide anions (O2) and thus mitigated cellular apoptosis. The time-course analysis of PA accumulation and ERK-like mitogen-activated protein kinase (MAPK) regulation indicated PA generation preceded MAPK activation, suggesting that the rapid accumulation of PA might be required for the initial MAPK activity. After 2 h of Ce4+ elicitation, however, PA-induced O2 burst, forming a negative regulation to MAPK activity, which in turn led to apoptotic signaling development.  相似文献   

19.
Animals with high metabolic rates are believed to have high rates of carbon and nitrogen isotopic incorporation. We hypothesized that (1) chronic exposure to cold, and hence an increase in metabolic rate, would increase the rate of isotopic incorporation of both 13C and 15N into red blood cells; and (2) that the rate of isotopic incorporation into red blood cells would be allometrically related to body mass. Two groups of sparrows were chronically exposed to either 5 or 22°C and switched from a 13C-depleted C3-plant diet to a more 13C-enriched C4-plant one. We used respirometry to estimate the resting metabolic rate of birds exposed chronically to our two experimental temperatures. The allometric relationship between the rate of 13C incorporation into blood and body mass was determined from published data. The of birds at 5°C was 1.9 times higher than that of birds at 22°C. Chronic exposure to a low temperature did not have an effect on the rate of isotopic incorporation of 15N save for a very small effect on the incorporation of 13C. The isotopic incorporation rate of 13C was 1.5 times faster than that of 15N. The fractional rate of 13C incorporation into avian blood was allometrically related to body mass with an exponent similar to −1/4. We conclude that the relationship between metabolic rate and the rate of isotopic incorporation into an animal’s tissues is indirect. It is probably mediated by protein turnover and thus more complex than previous studies have assumed.  相似文献   

20.
Others have shown that exposing oocytes to high levels of (10–20 mM) causes a paradoxical fall in intracellular pH (pHi), whereas low levels (e.g., 0.5 mM) cause little pHi change. Here we monitored pHi and extracellular surface pH (pHS) while exposing oocytes to 5 or 0.5 mM NH3/NH4 +. We confirm that 5 mM causes a paradoxical pHi fall (−ΔpHi ≅ 0.2), but also observe an abrupt pHS fall (−ΔpHS ≅ 0.2)—indicative of NH3 influx—followed by a slow decay. Reducing [NH3/NH4 +] to 0.5 mM minimizes pHi changes but maintains pHS changes at a reduced magnitude. Expressing AmtB (bacterial Rh homologue) exaggerates −ΔpHS at both levels. During removal of 0.5 or 5 mM NH3/NH4 +, failure of pHS to markedly overshoot bulk extracellular pH implies little NH3 efflux and, thus, little free cytosolic NH3/NH4 +. A new analysis of the effects of NH3 vs. NH4 + fluxes on pHS and pHi indicates that (a) NH3 rather than NH4 + fluxes dominate pHi and pHS changes and (b) oocytes dispose of most incoming NH3. NMR studies of oocytes exposed to 15N-labeled show no significant formation of glutamine but substantial accumulation in what is likely an acid intracellular compartment. In conclusion, parallel measurements of pHi and pHS demonstrate that NH3 flows across the plasma membrane and provide new insights into how a protein molecule in the plasma membrane—AmtB—enhances the flux of a gas across a biological membrane.
Walter F. Boron (Corresponding author)Email:
  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