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

2.
B. Liedvogel  R. Bäuerle 《Planta》1986,169(4):481-489
Chloroplasts from the cotyledons of mustard (Sinapis alba L.) seedlings were isolated on Percoll gradients, and showed a high degree of intactness (92%) and purity as judged by electron microscopy and marker-enzyme analysis (cytoplasmic contamination lower than 0.4% on a protein basis). The chloroplasts synthesized longchain fatty acids from both precursors [1-14C] acetate and [2-14C]pyruvate; maximum incorporation rates were 96 nmol·(mg Chl)-1·h-1 for acetate and 213 nmol·(mg Chl)-1·h-1 for pyruvate. Acetyl-CoA-producing enzymatic activities, namely acetyl-CoA synthetase (EC 6.2.1.1.) and a pyruvate dehydrogenase complex, showed specific activities of 14.8 nmol·(mg protein)-1·min-1 and 18.2 nmol·(mg protein)-1·min-1, respectively. The glycolytic enzymes phosphoglyceromutase (EC 2.7.5.3) phosphopyruvate hydratase (EC 4.2.1.11) and pyruvate kinase (EC 2.7.1.40) were all found to be components of these chloroplasts, thus indicating a possible pathway for intraplastid acetyl-CoA formation.Abbreviations ACS acetyl coenzyme A synthetase - Chl chlorophyll - DTE 1,4-dithioerythritol - PDHC pyruvate dehydrogenase complex - 3-PGA 3-phosphoglyceric acid  相似文献   

3.
Tobacco (Nicotiana tabacum L.) plants transformed with antisense rbcS to decrease the expression of ribulose-1,5-bisphosphate carboxylase-oxygenase (Rubisco) have been used to investigate the contribution of Rubisco to the control of photosynthesis in plants growing at different irradiances. Tobacco plants were grown in controlled-climate chambers under ambient CO2 at 20°C at 100, 300 and 750 mol·m–2·s–1 irradiance, and at 28°C at 100, 300 and 1000 mol·m–2·s–1 irradiance. (i) Measurement of photosynthesis under ambient conditions showed that the flux control coefficient of Rubisco (C infRubisco supA ) was very low (0.01–0.03) at low growth irradiance, and still fairly low (0.24–0.27) at higher irradiance. (ii) Short-term changes in the irradiance used to measure photosynthesis showed that C infRubisco supA increases as incident irradiance rises, (iii) When low-light (100 mol·m–2·s–1)-grown plants are exposed to high (750–1000 mol·m–2·s–1) irradiance, Rubisco is almost totally limiting for photosynthesis in wild types. However, when high-light-grown leaves (750–1000 mol·m–2·s–1) are suddenly exposed to high and saturating irradiance (1500–2000 mol·m–2·s–1), C infRubisco supA remained relatively low (0.23–0.33), showing that in saturating light Rubisco only exerts partial control over the light-saturated rate of photosynthesis in sun leaves; apparently additional factors are co-limiting photosynthetic performance, (iv) Growth of plants at high irradiance led to a small decrease in the percentage of total protein found in the insoluble (thylakoid fraction), and a decrease of chlorophyll, relative to protein or structural leaf dry weight. As a consequence of this change, high-irradiance-grown leaves illuminated at growth irradiance avoided an inbalance between the light reactions and Rubisco; this was shown by the low value of C infRubisco supA (see above) and by measurements showing that non-photochemical quenching was low, photochemical quenching high, and NADP-malate dehydrogenase activation was low at the growth irradiance. In contrast, when a leaf adapted to low irradiance was illuminated at a higher irradiance, Rubisco exerted more control, non-photochemical quenching was higher, photochemical quenching was lower, and NADP-malate dehydrogenase activation was higher than in a leaf which had grown at that irradiance. We conclude that changes in leaf composition allow the leaf to avoid a one-sided limitation by Rubisco and, hence, overexcitation and overreduction of the thylakoids in high-irradiance growth conditions, (v) Antisense plants with less Rubisco contained a higher content of insoluble (thylakoid) protein and chlorophyll, compared to total protein or structural leaf dry weight. They also showed a higher rate of photosynthesis than the wild type, when measured at an irradiance below that at which the plant had grown. We propose that N-allocation in low light is not optimal in tobacco and that genetic manipulation to decrease Rubisco may, in some circumstances, increase photosynthetic performance in low light.Abbreviations A rate of photosynthesis - C infRubisco supA flux control coefficient of Rubisco for photosynthesis - ci internal CO2 concentration - qE energy-dependent quenching of chlorophyll fluorescense - qQ photochemical quenching of chlorophyll fluorescence - NADP-MDH NADP-dependent malate dehydrogenase - Rubisco ribulose-1,5-bisphosphate carboxylase-oxygenase - RuBP ribulose-1,5-bisphosphate This work was supported by the Deutsche Forschungsgemeinschaft (SFB 137).  相似文献   

4.
Phosphoenolpyruvate carboxylase (PEPCase; EC 4.1.1.31) activity was found to be modulated by light and darkness when measured in the presence of K+, which had been added to induce swelling of guard-cell protoplasts (GCPs) from Vicia faba L., whereas no modulation was detected in the absence of K+ (PEPcase activity remained constant at 1.5±0.15 pmol PEP metabolized · GCP–1 ·h–1; subsequently, pmol GCP–1 ·h–1 will be used). The activity of PEPCase increased by 100% (from 1.5 to 3 pmol·protoplast–1·h–1) in darkness and by 200% (from 1.7 to 5 pmol·protoplast–1· h–1) in light and oscillations in activity of these magnitudes were repeated at intervals of 2 min (dark) and 2.5 min (light) for a period of 10 min during K+-induced increase in the volume of GCPs. The oscillations were reflected in changes in malate-pool sizes determined in plastids, mitochondria and the supernatant fraction (consisting of the cytosol and the vacuole). Malate probably functioned as a mitochondrial substrate, thus supplying ATP for K+ uptake and the swelling of the protoplasts. On the basis of the present paper and previous results (H. Schnabl and B. Michalke 1988, Life Sci. Adv. Plant Physiol. 7, 203–207) involving adenine nucleotidepool sizes in fractionated GCPs, a model is proposed to explain the cause-effect relationship between K+, PEPCase, the cytosolic and mitochondrial malate levels and ATP levels during the K+-induced increase of GCP volume.Abbreviations GCP dtguard-cell protoplast - PEP phosphoenol-pyruvate - PEPCase PEP carboxylase The authors thank Professor Hermann Schnabl, University of Stuttgart (FRG), for his assistance in applying the graph theory analysis. This work was supported by Deutsche Forschungsgemeinschaft to H.S.  相似文献   

5.
The suspension feeding of Bithynia tentaculata was tested in laboratory experiments. The animals were fed in 1-1 aerated glass beakers, and filtration rates were calculated from changes in cell concentrations during the 6-h experiment. Temperature influenced the filtering rate, with minimum values of 5ml · ind–1 · h–1 at 5° C and maxima of 17.2 ml · ind–1 · h–1 at 18° C. Three food species of different size, motility and cell surface characteristics (Chlamydomonas reinhardii, Chlorella vulgaris and Chlorogonium elongatum) did not affect filtration rates. Suspension feeding increased with increasing food concentrations up to 12 nl · ml–1, above which feeding rate was kept constant by lowering the filtering rates. Even the smallest animals tested (<4 mm body length) were found to be feeding on suspended food at a rate of 2.7 ml · ind–1 · h–1, and increasing rates up to 8.4 ml were found in the 6–7 mm size class. All size classes of Bithynia showed a circannual fluctuation of their filtration rates. The ecological consequences of Bithynia's ability to switch between two feeding modes, grazing and suspension feeding, are discussed.  相似文献   

6.
The growth of the anaerobic acetogenic bacterium Acetobacterium woodii DSM 1030 was investigated in fructose-limited chemostat cultures. A defined medium was developed which contained fructose, mineral salts, cysteine · HCl and Ca pantothenate (1 mg · 1–1) supplied in a vitamin supplement. Growth at high dilution rates was dependent on the presence of CO2 in the gas phase. The max was found to be 0.16 h–1 and the fructose maintenance requirement was 0.1 to 0.13 mmol fructose · (g dry wt)–1 · h–1. A growth yield of 61 g dry wt · (mol fructose)–1, corrected for the cell maintenance requirement and for incorporation of fructose carbon into cell biomass, was determined from the fructose consumption. A corresponding growth yield of 69 g dry wt · (mol fructose)–1 was calculated from the acetate production assuming that fructose fermentation was homoacetogenic. A YATP of 12.2 to 13.8 g dry wt · (mol ATP)–1 was calculated from these growth yields using a value of 5 mol ATP · (mol fructose)–1 as an estimate of the amount of ATP synthesised from fructose fermentation. The addition of yeast extract (0.5 g · 1–1) to the medium did not influence the max or cell yield. After prolonged growth under fructose-limited conditions the requirement of the culture for CO2 in the gas phase was reduced.Abbreviations YE yeast extract - IC inorganic carbon - D fermenter dilution rate : h–1 - MX maintenance requirement for X: mmol X · (g dry wt)–1 · h–1 - X may be fructose (Fruct), fructose consumed in energy metabolism (Fruct [E]), acetate (Ac) - ATP CO2, NH inf4 sup+ or Pi - qX specific rate of utilisation or consumption of X: mmol X · (g dry wt)–1 · h–1 - V fermenter volume: litre - rC · Cell, fermenter cell carbon production: mmol C · h–1 - YX yield of cells on X: g dry wt · (mol X)–1 - Y infx supmax the yield corrected for cell maintenance: g dry wt · (mol X)–1 - SATP stoichiometry of ATP synthesis from fructose: mol ATP · (mol frucose)–1 - x cell concentration: g dry wt · 1–1 - specific growth rate : h–1 - max maximum specific growth rate: h–1  相似文献   

7.
Summary Some environmental affects on cell aggregation described in the literature are briefly summarized. By means of a biomass recirculation culture (Contact system), using the yeast Torulopsis glabrata, the aggregation behavior of cells in static and in dynamic test systems is described. Sedimentation times required to obtain 50 g · l–1 yeast dry matter in static systems were always higher than in dynamic ones.In addition to, influencing the biomass yield, the specific growth rate of the yeast also affected cell aggregation. The specific growth rate and therefore the aggregation could be regulated by the biomass recirculation rate as well as by the sedimenter volume.Abbreviations fo Overflow flow rate (l·h–1) - fR Recycle flow rate (l·h–1) - ft0t Total flow rate through the fermenter (l·h–1) - g Gram - h Hour - DR Fermenter dilution rate due to recycle (h–1) - DS Fermeter dilution rate due to substrate (h–1) - Dtot Total fermenter dilution rate (h–1) - l Liter - Specific growth rate (h–1) - PF Fermenter productivity (g·l–1·h–1) - PFS Overall productivity (g·l–1·h–1) - RpM Rates per minute - RS Residual sugar content in the effluent with respect to the substrate concentration (%) - Y Yield of biomass with respect to sugar concentration (%) - Sed 50 Sedimentation time to reach a YDM of 50 g·l–1 (min) - V Volume (l) - VF Fermenter volume (l) - VSed Sedimenter volume (l) - VVM Volumes per volume and minute - XF YDM in the fermenter (g·l–1) - XF YDM in the recycle (g·l–1) - XS Yeast dry matter due to substrate concentration (g·l–1) - YDM Yeast dry matter (g·l–1)  相似文献   

8.
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   

9.
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   

10.
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   

11.
The photosynthetic properties of a range of lichens (eight species) containing green algal primary photobionts of either the genus Coccomyxa, Dictyochloropsis or Trebouxia were examined with the aim of obtaining a better understanding for the different CO2 acquisition strategies of lichenized green algae. Fast transients of light/dark-dependent CO2 uptake and release were measured in order to screen for the presence or absence of a photosynthetic CO2-concentrating mechanism (CCM) within the photobiont. It was found that lichens with Trebouxia photobionts (four species) were able to accumulate a small pool of inorganic carbon (DIC; 70–140 nmol per mg chlorophyll (Chl)), in the light, which theoretically may result in, at least, a two to threefold increase in the stromal CO2 concentration, as compared to that in equilibrium with ambient air. The other lichens (four species), which were tripartite associations between a fungus, a cyanobacterium (Nostoc) and a green alga (Coccomyxa or Dictyochloropsis) accumulated a much smaller pool of DIC (10–30 nmol·(mg Chl)–1). This pool is most probably associated with the previously documented CCM of Nostoc, inferred from the finding that free-living cells of Coccomyxa did not show any signs of DIC accumulation. In addition, the kinetics of fast CO2 exchange for free-living Nostoc were similar to those of intact tripartite lichens, especially in their responses to the CCM and the carbonic anhydrase (CA) inhibitor ethoxyzolamide. Trebouxia lichens had a higher photosynthetic capacity at low and limiting external CO2 concentrations, with an initial slope of the CO2-response curve of 2.6–3.9 mol·(mg Chl)–1·h–1·Pa–1, compared to the tripartite lichens which had an initial slope of 0.5–1.1 mol-(mg Chl)–1·h–1·-Pa–1, suggesting that the presence of a CCM in the photobiont affects the photosynthetic performance of the whole lichen. Regardless of these indications for the presence or absence of a CCM, ethoxyzolamide inhibited the steady-state rate of photosynthesis at low CO2 in all lichens, indicating a role of CA in the photosynthetic process within all of the photobionts. Measurements of CA activity in photobiont-enriched homogenates of the lichens showed that Coccomyxa had by far the highest activity, while the other photobionts displayed only traces or no activity at all. As the CCM is apparently absent in Coccomyxa, it is speculated that this alga compensates for this absence with high internal CA activity, which may function to reduce the CO2-diffusion resistance through the cell.Abbreviations CA carbonic anhydrase (EC 4.2.1.1) - CCM CO2-concentrating mechanism - Chl chlorophyll - DIC dissolved inorganic carbon - EZ ethoxyzolamide or 6-ethoxy-2-benzo-thiazole-2-sulfonamide - GA glycolaldehyde - Hepps 4-(2-hydroxyethyl)-l-piperazinepropanesulfonic acid - Rubisco ribulose-1,5-bisphosphate carboxylase-oxygenase (EC 4.1.1.39) This research was supported by a grant from the Swedish Natural Sciences Resource Council to K.P.  相似文献   

12.
Rhodopseudomonas acidophila strain 7050 can satisfy all its nitrogen and carbon requirements from l-alanine. Addition of 100 M methionine sulfoximine to alanine grown cultures had no effect on growth rate indicating that deamination of alanine via alanine dehydrogenase and re-assimilation of the released NH 4 + by glutamine synthetase/glutamate synthase was an insignificant route of nitrogen transfer in this bacterium. Determination of aminotransferase activities in cell-free extracts failed to demonstrate the presence of direct routes from alanine to either aspartate or glutamate. The only active aminotransferase involving l-alanine was the alanine-glyoxylate enzyme (114–167 nmol·min–1·mg–1 protein) which produced glycine as end-product. The amino group of glycine was further transaminated to yield aspartate via a glycineoxaloacetate aminotransferase (117–136 nmol·min–1 ·mg–1 protein). No activity was observed when 2-oxoglutarate was substituted for oxaloacetate. The formation of glutamate from aspartate was catalysed by aspartate-2-oxoglutarate aminotransferase (85–107 nmol·min–1·mg–1 protein). Determinations of free intracellular amino acid pools in alanine and alanine+100 M methionine sulfoximine grown cells showed the predominance of glutamate, glycine and aspartate, providing further evidence that in alanine grown cultures R. acidophila satisfies its nitrogen requirements for balanced growth by transamination.Abbreviations ADH alanine dehydrogenase - GDH glutamate dehydrogenase - GS glutamine synthetase - GOGAT glutamate synthase - MSO methionine sulfoximine - GOT glutamate-oxaloacetate aminotransferase - GPT glutamate-pyruvate amino-transferase - AGAT alanine-glyoxylate aminotransferase - GOAT glycine-oxaloacetate aminotransferase - GOTAT glycine-2-oxoglutarate aminotransferase - AOAT alanine-oxaloacetate aminotransferase  相似文献   

13.
Summary Submerged batch cultivation under controlled environmental conditions of pH 3.8, temperature 30°C, and KLa200 h–1 (above 180 mMO2 l –1 h–1 oxygen supply rate) produced a maximum (12.0 g·l –1) SCP (Candida utilis) yield on the deseeded nopal fruit juice medium containing C/N ratio of 7.0 (initial sugar concentration 25 g·l –1) with a yield coefficient of 0.52 g cells/g sugar. In continuous cultivation, 19.9 g·l –1 cell mass could be obtained at a dilution rate (D) of 0.36 h–1 under identical environmental conditions, showing a productivity of 7.2 g·l –1·h–1. This corresponded to a gain of 9.0 in productivity in continuous culture over batch culture. Starting with steady state values of state variables, cell mass (CX–19.9 g·l –1), limiting nutrient concentration (Cln–2.5 g·l –1) and sugar concentration (CS–1.5 g·l –1) at control variable conditions of pH 3.8, 30°C, and KLa 200 h–1 keeping D=0.36 h–1 as reference, transient response studies by step changes of these control variables also showed that this pH, temperature and KLa conditions are most suitable for SCP cultivation on nopal fruit juice. Kinetic equations obtained from experimental data were analysed and kinetic parameters determined graphically. Results of SCP production from nopal fruit juice are described.Nomenclature Cln concentration of ammonium sulfate (g·l –1) - CS concentration of total sugar (g·l –1) - CX cell concentration (g·l –1) - D dilution rate (h–1) - Kln Monod's constant (g·l –1) - m maintenance coefficient (g ammonium sulfate cell–1 h–1) - m(S) maintenance coefficient (g sugar g cell–1 h–1) - t time, h - Y yield coefficient (g cells/g ammonium sulfate) - Ym maximum of Y - YS yield coefficient based on sugar consumed (g cells · g sugar–1) - YS(m) maximum value of YS - µm maximum specific growth rate constant (h–1)  相似文献   

14.
Michael A. Grusak 《Planta》1995,197(1):111-117
To understand the whole-plant processes which influence the Fe nutrition of developing seeds, we have characterized root Fe(III)-reductase activity and quantified whole-plant Fe balance throughout the complete 10-week (10-wk) life cycle of pea (Pisum sativum L., cv. Sparkle). Plants were grown hydroponically in complete nutrient solution with a continuous supply of chelated Fe; all side shoots were removed at first appearance to yield plants with one main shoot. Root Fe(III)-reductase activity was assayed with Fe(III)-EDTA. Flowering of the experimental plants began on wk 4 and continued until wk 6; seed growth and active seed import occurred during wks 5–10. Vegetative growth terminated at wk 6. Iron(III) reduction in whole-root systems was found to be dynamically modulated throughout the plant's life cycle, even though the plants were maintained on an Fe source. Iron(III)-reductase activity ranged from 1–3 mol Fe reduced · g –1 DW · h–1 at early and late stages of the life cycle to 9.5 mol Fe reduced · g–1 DW · h–1 at wk 6. Visual assays demonstrated that Fe(III)-reductase activity was localized to extensive regions of secondary and tertiary lateral roots during this peak activity. At midstages of growth (wks 6–7), root Fe(III)-reductase activity could be altered by changes in internal shoot Fe demand or external root Fe supply: removal of all pods or interruption of phloem transport from the reproductive portion of the shoot (to the roots) resulted in lowered root Fe(III)-reductase activity, while removal of Fe from the nutrient solution resulted in a stimulation of this activity. Total shoot Fe content increased throughout the 10-wk growth period, with Fe content in the non-seed tissues of the shoot declining by 50% of their maximal level and accounting for 35% of final seed Fe content. At maturity, total seed Fe represented 74% of total shoot Fe; total Fe in the roots (apoplasmic and symplasmic Fe combined) was minimal. These studies demonstrate that the root Fe(III)-reductase system responds to Fe status and/or Fe requirements of the shoot, apparently through shoot-to-root communication involving a phloem-mobile signal. During active seed-fill, enhanced root Fe(III)-reductase activity is necessary to generate sufficient Fe2+ for continued root Fe acquisition. This continuing Fe supply to the shoot is essential for the developing seeds to attain their Fe-content potential. Increased rates of root Fe(III) reduction would be necessary for seed Fe content to be enhanced in Pisum sativum.Abbreviations BPDS bathophenanthrolinedisulfonic acid - DAF days after flowering - DW dry weight - EDDHA N,N-ethylenebis[2-(2-hydroxyphenyl)-glycine] - wk week This project has been funded in part with federal funds from the U.S. Department of Agriculture, Agricultural Research Service under Cooperative Agreement number 58-6250-1-003. The contents of this publication do not necessarily reflect the views or policies of the U.S. Department of Agriculture, nor does mention of trade names, commercial products, or organizations imply endorsement by the U.S. Government. The author wishes to acknowledge S. Pezeshgi and K. Koch for their excellent technical assistance, L. Loddeke for editorial comments, and A. Gillum for assistance with the figures.  相似文献   

15.
Clearance rates of sessile rotifers: in vitro determinations   总被引:1,自引:1,他引:0  
We measured laboratory clearance rates of 10 rotifer and one unidentified bryozoan species from 3 different lakes using 32P labeled algae (Chlamydomonas) or yeast (Rhodotorula). Clearance rates for all rotifers fed yeast ranged from < 2.0 to > 260 µl · animal–1 · h–1 depending on species. The in vitro clearance rates of two sessile rotifers (Ptygura crystallina and P. pilula) were not significantly different from previously measured in situ rates (Wallace and Starkweather 1983). Clearance rates for 5 rotifers fed algae ranged from < 5.0 to > 90.0 µl · animal–1 · h–1. Ptygura beauchampi, P. crystallina, P. pilula, Floscularia conifera, and F. melicerta ingested both cell types but their clearance rates varied substantially among species and between cell types. There was a substantial time-dependent loss of 32P from formalin-fixed animals (Sinantherina socialis) awaiting processing. This loss stabilized at approximately 20 hours and was estimated to be about 40% of the initial ingested label. Clearance rates for the bryozoan fed yeast or algae were highly variable, ranging from < 1.0 to > 3 000 µl · animal–1 · h–1.  相似文献   

16.
Acclimation to changes in the light environment was investigated in Arabidopsis thaliana (L.) Heynh. cv. Landsberg erecta. Plants grown under four light regimes showed differences in their development, morphology, photosynthetic performance and in the composition of the photosynthetic apparatus. Plants grown under high light showed higher maximum rates of oxygen evolution and lower levels of light-harvesting complexes than their low light-grown counterparts; plants transferred to low light showed rapid changes in maximum photosynthetic rate and chlorophyll-a/b ratio as they became acclimated to the new environment. In contrast, plants grown under lights of differing spectral quality showed significant differences in the ratio of photosystem II to photosystem I. These changes are consistent with a model in which photosynthetic metabolism provides signals which regulate the composition of the thylakoid membrane.Abbreviations Aac1 gene encoding actin - Chl chlorophyll - F far-red-enriched light (R:FR = 0.72) - FR far-red light - H high light (400 mol · m–2 · s–1) - L low light (100 ml · m–2 · s–1) - LHCII light-harvesting complex of PSII - Lhcb genes encoding the proteins of LHCII - R red light - Rbcs genes encoding the small subunit of Rubisco - Rubisco ribulose-1,5-bisphosphate carboxylase/oxygenase - W white light (R:FR = 1.40) This work was supported by Natural Environment Research Council Grant No. GR3/7571A. We would like to thank H. Smith (Botany Department, University of Leicester) and E. Murchie (University of Sheffield) for helpful discussions.  相似文献   

17.
Long-term chilling of young tomato plants under low light   总被引:8,自引:0,他引:8  
The properties of two Calvin-cycle key enzymes, i.e. stromal fructose-1,6-bisphosphatase (sFBPase) and ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) were studied in the cultivated tomato (Lycopersicon esculentum Mill.) and in four lines of a wild tomato (L. peruvianum Mill.) from different altitudes. During chilling for 14 d at 10°C and low light, the activation energy (EA) of the reaction catalyzed by sFBPase decreased by 5–10 kJ·mol–1 inL. esculentum and the threeL. peruvianum lines from high altitudes. InL. peruvianum, no loss or only small losses of enzyme activity were observed during the chilling. Together with the change in EA, this indicates that the latter species is able to acclimate its Calvin-cycle enzymes to low temperatures. InL. esculentum, the chilling stress resulted in the irreversible loss of 57% of the initial sFBPase activity. Under moderately photoinhibiting chilling conditions for 3 d, theL. peruvianum line from an intermediate altitude showed the largest decreases in both the ratio of variable to maximum chlorophyll fluorescence (Fv/Fm) and the in-vivo activation state of sFBPase, while the otherL. peruvianum lines showed no inhibition of sFBPase activation. Ribulose-1,5-bisphosphate carboxylase/oxygenase was isolated by differential ammonium-sulfate precipitation and gel filtration and characterized by two-dimensional electrophoresis. The enzyme fromL. esculentum had three isoforms of the small subunit of Rubisco, each with different isoelectric points. Of these, theL. peruvianum enzyme contained only the two more-acidic isoforms. Arrhenius plots of the specific activity of purified Rubisco showed breakpoints at approx. 17°C. Upon chilling, the specific activity of the enzyme fromL. esculentum decreased by 51%, while EA below the breakpoint temperature increased from 129 to 189 kJ·mol–1. In contrast, Rubisco from theL. peruvianum lines from high altitudes was unaffected by chilling. We tested several possibile explanations for Rubisco inactivation, using two-dimensional electrophoresis, analytical ultracentrifugation, gel filtration and inhibitor tests. No indications were found for differential expression of the subunit isoforms, proteolysis, aggregation, subunit disassembly, or inhibitor accumulation in the enzyme from chilledL. esculentum. We suggest that the activity loss in theL. esculentum enzyme upon chilling is the result of a modification of sulfhydryl groups or other sidechains of the protein.Abbreviations a.s.l. above sea level - Chl chlorophyll - DTT dithiothreitol - EA activation energy - FBP fructose-1,6-bisphosphate - Fv/Fm ratio of variable to maximum chlorophyll fluorescence - HL high light (500 mol photons·m–2·s–1) - LSU large subunit of Rubisco - ME 2-mercaptoethanol - Rubisco ribulose-1,5-bisphosphate carboxylase/oxygenase - RuBP ribulose-1,5-bisphosphate - sFBPase stromal fructose-1,6-bisphosphatase - SSU small subunit of Rubisco  相似文献   

18.
Experiments were carried out to determine how decreased expression of ribulose-1,5-bisphosphate carboxylase-oxygenase (Rubisco) affects photosynthetic metabolism in ambient growth conditions. In a series of tobacco (Nicotiana tabacum L.) plants containing progressively smaller amounts of Rubisco the rate of photosynthesis was measured under conditions similar to those in which the plants had been grown (310 mol photons · m–2 · s–1, 350 bar CO2, 22° C). (i) There was only a marginal inhibition (6%) of photosynthesis when Rubisco was decreased to about 60% of the amount in the wildtype. The reduced amount of Rubisco was compensated for by an increase in Rubisco activation (rising from 60 to 100%), with minor contributions from an increase of its substrates (ribulose-1,5-bisphosphate and the internal CO2 concentration) and a decrease of its product (glycerate-3-phosphate). (ii) The decreased amount of Rubisco was accompanied by an increased ATP/ADP ratio that may be causally linked to the increased activation of Rubisco. An increase of highenergy-state chlorophyll fluorescence shows that thylakoid membrane energisation and high-energy-state-dependent energy dissipation at photosystem two had also increased. (iii) A further decrease of Rubisco (in the range of 50–20% of the wildtype level) resulted in a strong and proportional inhibition of CO2 assimilation. This was accompanied by a decrease of fructose-1,6-bisphosphatase activity, coupling-factor 1 (CF1)-ATP-synthase protein, NADP-malate dehydrogenase protein, and chlorophyll. The chlorophyll a/b ratio did not change, and enolase and sucrose-phosphate synthase activity did not decrease. It is argued that other photosynthetic enzymes are also decreased once Rubisco decreases to the point at which it becomes strongly limiting for photosynthesis. (iv) It is proposed that the amount of Rubisco in the wildtype represents a balance between the demands of light, water and nitrogen utilisation. The wildtype overinvests about 15% more protein in Rubisco than is needed to avoid a strict Rubisco limitation of photosynthesis. However, this excess Rubisco allows the wildtype to operate with lower thylakoid energisation, and decreased high-energy-state-dependent energy dissipation, hence increasing light-use efficiency by about 6%. It also allows the wildtype to operate with a lower internal CO2 concentration in the leaf and a lower stomatal conductance at a given rate of photosynthesis, so that instantaneous water-use efficiency is marginally (8%) increased.Abbreviations Ci CO2 concentration in the air spaces within the leaf - CF1 coupling factor 1 - Chl chlorophyll Fru1 - 6bisP fructose-1,6-bisphosphate - Fm fluorescence yield with a saturating pulse in dark-adapted material - Fo ground-level of fluorescence obtained using a weak non-actinic modulated beam in the dark - PGA glycerate-3-phosphate - rbcS gene for the nuclear-encoded small subunit of Rubisco - Rubisco ribulose-1,5-bisphosphate carboxylase-oxygenase - Ru1, 5bisP ribulose-1,5-bisphosphate  相似文献   

19.
Summary A detailed study on the reductive amination of -ketoisovalerate to l-valine by l-valine dehydrogenase using glucose dehydrogenase as an NADH regeneration enzyme was performed. The presence of both enzyme activities in Bacillus megaterium ATCC 39 118 permitted a direct and systematic comparison of the performances (initial l-valine production rate, productivity, molar conversion yield) of different types of conversion systems: purified enzymes or crude extract and whole cells, intact or permeabilized. A maximal l-valine productivity of 8 mmol·l–1 · h–1 was obtained using purified enzymes which constituted the most efficient system with a maximal rate of 0.87 mol · ml–1 · min–1 and a molar conversion yield of 0.91. Permeabilized cells were also an attractive system because of their easy preparation and of the good performances attained.Offprint requests to: F. Monot  相似文献   

20.
Homogeneous populations of developing microspores and pollen from anthers of lily (Lilium longiflorum Thumb.) and tobacco (Nicotiana tabacum L.) show a continuous production of biomass, reaching a maximum in young pollen. The rate of RNA synthesis was 460 fg · h–1 in young binucleate cells, 138 fg · h–1 in late binucleate cells and 56 fg · h–1 in microspores. The mRNA population in developing pollen can be separated into three groups. In the first group, certain types of mRNAs are present at a constant level during all stages of development. A second group is characteristic of young pollen and increases quantitatively until anthesis. A third group is seen transiently; to this belong mRNAs present only before mitosis or at a distinct cell stage after mitosis. Some of the translation products of this latter group of mRNAs showed similarities between lily and tobacco on two-dimensional gels in respect of molecular weight and isolectric point, indicating that those mRNAs and proteins play a role in the regulation of pollen development.Abbreviations cDNA copy DNA - pI isolectric point To whom correspondence should be addressed.  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