首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 484 毫秒
1.
Thylakoids isolated from cells of the red alga Porphyridium cruentum exhibit an increased PS I activity on a chlorophyll basis with increasing growth irradiance, even though the stoichiometry of Photosystems I and II in such cells shows little change (Cunningham et al. (1989) Plant Physiol 91: 1179–1187). PS I activity was 26% greater in thylakoids of cells acclimated at 280 mol photons · m–2 · s–1 (VHL) than in cells acclimated at 10 mol photons · m–2 · s–1 (LL), indicating a change in the light absorbance capacity of PS I. Upon isolating PS I holocomplexes from VHL cells it was found that they contained 132±9 Chl/P700 while those obtained from LL cells had 165±4 Chl/P700. Examination of the polypeptide composition of PS I holocomplexes on SDS-PAGE showed a notable decrease of three polypeptides (19.5, 21.0 and 22 kDa) in VHL-complexes relative to LL-complexes. These polypeptides belong to a novel LHC I complex, recently discovered in red algae (Wolfe et al. (1994a) Nature 367: 566–568), that lacks Chl b and includes at least six different polypeptides. We suggest that the decrease in PS I Chl antenna size observed with increasing irradiance is attributable to changes occurring in the LHC I-antenna complex. Evidence for a Chl-binding antenna complex associated with PS II core complexes is lacking at this point. LHC II-type polypeptides were not observed in functionally active PS II preparations (Wolfe et al. (1994b) Biochimica Biophysica Acta 1188: 357–366), nor did we detect polypeptides that showed immunocross-reactivity with LHC II specific antisera (made to Chlamydomonas and Euglena LHC II).Abbreviations Bis-Tris bis(2-hydroxyethyl)imino-tris(hydroxymethyl)methane - DCPIP 2,6-dichlorophenol indophenol - -dm dodecyl--d-maltoside - HL high light of 150 mol photons · m–2 · s–1 - LGB lower green band - LHC I light-harvesting complex of PS I - LHC II light-harvesting complex of PS II - LL low light of 10 mol photons · m–2 · s–1 - ML medium light of 50 mol photons · m–2 · s–1 - MES 2-(N-morpholino) ethanesulfonic acid - P700 reaction center of PS I - PFD photon flux density - Trizma tris(hydroxymethyl)aminomethane - UGB upper green band - VHL very high light of 280 mol photons · m–2 · s–1  相似文献   

2.
Using primary cultures of gill pavement cells from freshwater rainbow trout, a method is described for achieving confluent monolayers of the cells on glass coverslips. A continuous record of intracellular pH was obtained by loading the cells with the pH-sensitive flourescent dye 2,7-bis(2-carboxyethyl)-5(6)-carboxyfluorescein and mounting the coverslips in the flowthrough cuvette of a spectrofluorimeter. Experiments were performed in HEPES-buffered media nominally free of HCO3. Resting intracellular pH (7.43 at extracellular pH=7.70) was insensitive to the removal of Cl or the application of 4-acetamido-4-isothiocyanatostilbene-2,2-disulfonic acid (0.1 mmol·l–1), but fell by about 0.3 units when Na+ was removed or in the presence of amiloride (0.2 mmol·l–1). Exposure to elevated ammonia (ammonia prepulse; 30 mmol·l–1 as NH4Cl for 6–9 min) produced an increase in intracellular pH (to about 8.1) followed by a slow decay, and washout of the pulse caused intracellular pH to fall to about 6.5. Intracellular non-HCO 3 buffer capacity was about 13.4 slykes. Rapid recovery of intracellular pH from intracellular acidosis induced by ammonia prepulse was inhibited more than 80% in Na+-free conditions or in the presence of amiloride (0.2 mmol·l–1). Neither bafilomycin A1 (3 mol·l–1) nor Cl removal altered the intracellular pH recovery rate. The K m for Na+ of the intracellular pH recovery mechanism was 8.3 mmol·l–1, and the rate constant at V max was 0.008·s–1 (equivalent to 5.60 mmol H+·l–1 cell water·min–1), which was achieved at external Na+ levels from 25 to 140 mmol·l–1. We conclude that intracellular pH in cultured gill pavement cells in HEPES-buffered, HCO 3 -free media, both at rest and during acidosis, is regulated by a Na+/H+ antiport and not by anion-dependent mechanisms or a vacuolar H+-ATPase.Abbreviations BCECF 2,7-bis(2-carboxyethyl)-5(6)-carboxy-fluorescein - BCECF/AM 2,7-bis(2-carboxyethyl)-5(6)-carboxy-fluorescein, acetoxymethylester - Cholin-Cl choline chloride - DMSO dimethyl sulfoxide - EDTA ethylene diamine tetra-acetic acid - FBS foetal bovine serum - H + -ATPase Proton-dependent adenosine triphosphatase - HEPES N-[2-hydroxyethyl]piperazine-N[2-ethanesulfonic acid] - pH i intracellular pH - pH e extracellular pH - PBS phosphate-buffered saline - SITS 4-acetamido-4-isothiocyanatostilbene-2,2-disulfonic acid  相似文献   

3.
Isolated embryos ofKarwinskia humboldtiana were cultured in vitro. The growth of embryos and development to plantlets on woody plant medium supplemented with indole-3-acetic acid 6.10-2 mol l–1, gibberellic acid (GA3) 3.10-2 mol l–1, and 6-benzylaminopurine (BA) 2 mol l–1 was obtained. Multiplication of shoots and rooting of excised shoots has been achieved. Callus formation on modified Murashige-Skoog medium supplemented with 1-naphthaleneacetic acid 10 mol l–1, GA3 14 mol l–1, and kinetin 5 mol l–1 on hypocotyls, or on root cultures on medium supplemented with 2.4-dichlorophenoxyacetic acid 10 mol l–1 and BA 10 mol l–1 was induced.Abbreviations BA 6-benzylaminopurine - 2,4-d 2,4-dichlorophenoxyacetic acid - GA3 gibberellic acid - IAA indole-3-acetic acid - NAA 1-naphthaleneacetic acid - TEM transmission electron microscopy  相似文献   

4.
Eicosapentaenoic (EPA) and docosahexaenoic (DHA) acid productivities from chemostat cultures of an isolate of Isochrysis galbana have been studied. The productivities reached in the interval of dilution rates between 0.0295 h–1 and 0.0355 h–1 were 1.5mg·1–1·h–1 for lipids, 300 g·1–1·h–1 for EPA and 130g1·1–1·h–1 for DHA. Furthermore, light attenuation by mutual shading, and agitation speed influences on growth and fatty acid composition were analysed. A model relating steady-state dilution rates to internal average light intensity has been proposed, the parameter values of which obtained by non-linear regression were: maximum specific growth rate (max)=0.0426 h–1; the affinity of cells to light (Ik) = 10.92 W·m–2; the exponent (n) = 5.13; regression coefficient (r 2)=0.9999. Correspondence to: E. Molina Grima  相似文献   

5.
The ecology of Lake Nakuru (Kenya)   总被引:11,自引:0,他引:11  
E. Vareschi 《Oecologia》1982,55(1):81-101
Summary Abiotic factors, standing crop and photosynthetic production were studied in the equatorial alkaline-saline closed-basin Lake Nakuru (cond. 10,000–160,000 S). Meteorological conditions and abiotic factors offer suppositions for a high primary productivity: mean solar radiation is 450–550 kerg·cm-2·s-1, with little seasonal variation, regular winds circulate the lake every day and nutrient concentrations are usually high (>100 g P–PO4·l-1). Oxygen concentrations near sediments were <1 gO2·m-3 for at least 6 h·d-1 in 1972/73, resulting in a release of 45 mg P–PO4·m-2·d-1. Attenuation coefficients vary from 3.6–16.5 according to algal densities and mean depth from 0–400 cm. Algal biomass was 200 g·m-3 (d.w.) in 1972/73, due to a lasting Spirulina platensis bloom (98.5% of algal biomass). In 1974 algal biomass suddenly dropped to 50 g·m-3 (d.w.). Spirulina and several consumer organisms almost vanished, but coccoid cyanobacteria, Anabaenopsis and diatoms increased. Several causes for this change in ecosystem structure are discussed. The use of the light/dark bottle method to measure photosynthetic production in eutrophic alkaline lakes is discussed and relevant experiments were done. Oxygen tensions of 2–35 gO2·m-3 do not influence primary production rates. Net photosynthetic rates (mgO2·m-3·h-1; photosynthetic quotient=1.18) reached 12–17.7 in 1972/73 and 2–3 in 1974, but vertically integrated rates were only 1–1.4 in 1972/73 and 0.8 in 1974, and daily net photosynthetic rates (gO2·m-3·24 h-1) 3.5 in 1972/73 and 1 in 1974. 50% of areal rates were produced within the 10 most productive cm of the depth profile. The disproportion between high algal standing crops and relatively low production rates is due to self-shading of the algae, reducing the euphotic zone to 35 cm in 1972/73 and 77 cm in 1974. Efficiency of light utilization is 0.4–2%, varying with time of day and phytoplankton density. In situ efficiencies show an inverse relationship to light intensities. Photosynthetic rates of L. Nakuru remain within the range of other African lakes (0.1–3 gO2·m-2·h-1). The relation of O2 produced/Chl a of the euphotic zone is 50% lower then in tropical African freshwater lakes and conforms to lakes of temperate regions.  相似文献   

6.
The obligate shade plant, Tradescantia albiflora Kunth grown at 50 mol photons · m–2 s–1 and Pisum sativum L. acclimated to two photon fluence rates, 50 and 300 mol · m–2 · s–1, were exposed to photoinhibitory light conditions of 1700 mol · m–2 · s–1 for 4 h at 22° C. Photosynthesis was assayed by measurement of CO2-saturated O2 evolution, and photosystem II (PSII) was assayed using modulated chlorophyll fluorescence and flash-yield determinations of functional reaction centres. Tradescantia was most sensitive to photoinhibition, while pea grown at 300 mol · m–2 · s–1 was most resistant, with pea grown at 50 mol · m–2 · s–1 showing an intermediate sensitivity. A very good correlation was found between the decrease of functional PSII reaction centres and both the inhibition of photosynthesis and PSII photochemistry. Photoinhibition caused a decline in the maximum quantum yield for PSII electron transport as determined by the product of photochemical quenching (qp) and the yield of open PSII reaction centres as given by the steady-state fluorescence ratio, FvFm, according to Genty et al. (1989, Biochim. Biophys. Acta 990, 81–92). The decrease in the quantum yield for PSII electron transport was fully accounted for by a decrease in FvFm, since qp at a given photon fluence rate was similar for photoinhibited and noninhibited plants. Under lightsaturating conditions, the quantum yield of PSII electron transport was similar in photoinhibited and noninhibited plants. The data give support for the view that photoinhibition of the reaction centres of PSII represents a stable, long-term, down-regulation of photochemistry, which occurs in plants under sustained high-light conditions, and replaces part of the regulation usually exerted by the transthylakoid pH gradient. Furthermore, by investigating the susceptibility of differently lightacclimated sun and shade species to photoinhibition in relation to qp, i.e. the fraction of open-to-closed PSII reaction centres, we also show that irrespective of light acclimation, plants become susceptible to photoinhibition when the majority of their PSII reaction centres are still open (i.e. primary quinone acceptor oxidized). Photoinhibition appears to be an unavoidable consequence of PSII function when light causes sustained closure of more than 40% of PSII reaction centres.Abbreviations Fo and Fo minimal fluorescence when all PSII reaction centres are open in darkness and steady-state light, respectively - Fm and Fm maximal fluorescence when all PSII reaction centres are closed in darkand light-acclimated leaves, respectively - Fv variable fluorescence - (Fm-Fo) under steady-state light con-ditions - Fs steady-state fluorescence in light - QA the primary,stable quinone acceptor of PSII - qNe non-photochemical quench-ing of fluorescence due to high energy state - (pH); qNi non-photochemical quenching of fluorescence due to photoinhibition - qp photochemical quenching of fluorescence To whom correspondence should be addressedThis work was supported by the Swedish Natural Science Research Council (G.Ö.) and the award of a National Research Fellowship to J.M.A and W.S.C. We thank Dr. Paul Kriedemann, Division of Forestry and Forest Products, CSIRO, Canberra, Australia, for helpful discussions.  相似文献   

7.
Summary The following equations represent the influence of the ethanol concentration (E) on the specific growth rate of the yeast cells () and on the specific production rate of ethanol () during the reactor filling phase in fed-batch fermentation of sugar-cane blackstrap molasses: = 0 - k · E and v = v 0 · K/(K +E) Nomenclature E ethanol concentration in the aqueous phase of the fermenting medium (g.L–1) - Em value of E when = 0 or = 0 (g.L–1) - F medium feeding rate (L.h–1) - k empirical constant (L.g–1.h–1) - K empirical constant (g.L–1) - Mas mass of TRS added to the, reactor (g) - Mcs mass of consumed TRS (g) - Me mass of ethanol in the aqueous phase of the fermenting medium (g) - Ms mass of TRS in the aqueous phase of the fermenting medium (g) - Mx mass of yeast cells (dry matter) in the fermenting medium (g) - r correlation coefficient - S TRS concentration in the aqueous phase of the fermenting medium (g.L–1) - Sm TRS concentration of the feeding medium (g.L–1) - t time (h) - T temperature (° C) - TRS total reducing sugars calculated as glucose - V volume of the fermenting medium (L) - V0 volume of the inoculum (L) - X yeast cells concentration (dry matter) in the fermenting medium (g.L–1) - filling-up time (h) - specific growth rate of the yeast cells (h–1) - 0 value of when E=0 - specific production rate of ethanol (h–1) - 0 value of when E=0 - density of the yeast cells (g.L–1) - dry matter content of the yeast cells  相似文献   

8.
Summary Transformation of cucumber cv. Endeavor was attempted using three Agrobacterium tumefaciens strains (a supervirulent leucinopine type, an octopine type and a nopaline type), each harbouring one of three binary vectors which contained an acidic chitinase gene from petunia, and basic chitinase genes from tobacco and bean, respectively, driven by the CaMV 35S promoter. Petiole explants were inoculated with a bacterial suspension (108 cells·ml–1), cocultivated for 48–96 h and placed on Murashige and Skoog (MS) medium with 5.0 M each of 2,4-D and BA, 50 mg·l–1 kanamycin and 500 mg·l–1 carbenicillin. The frequency of embryogenic callus formation ranged from 0 to 12%, depending on strains/vectors used and length of cocultivation, with the highest being obtained using the leucinopine strain with petunia acidic chitinase gene. The kanamycin-resistant embryogenic calli were used to initiate suspension cultures (in liquid MS medium with 1.0/1.0 M 2,4-D/BA, 50 mg·l–1 kanamycin) for multiplication of embryogenic cell aggregates. Upon plating of cell aggregates onto solid MS medium with 1.0/1.0 M NAA/BA and 50 mg·l–1 kanamycin, calli continued to grow and later differentiated into plantlets. Transformation by the leucinopine strain and all three vectors was confirmed by PCR amplification of the NPT II gene in transgenic calli and plants, in addition to Southern analysis. Expression of the acidic chitinase gene (from petunia) and both basic chitinase genes (from tobacco and bean) in different transgenic cucumber lines was confirmed by Western analyses.Abbreviations 2,4-D 2,4-dichlorophenoxyacetic acid - BA 6-benzyl-aminopurine - CaMV cauliflower mosaic virus - NAA naphthaleneacetic acid - NPT II neomycin phosphotransferase II - PCR polymerase chain reaction  相似文献   

9.
Schizosaccharomyces yeasts can be used for deacidification of grape musts. To this aim, we studied malic acid degradation by yeasts included in double layer alginate beads in a bubble column reactor. Use of immobilized micro-organisms allowed a continuous process with high dilution rates giving a deacidification capacity of 0.032 g of malate/hour/dm3/g of beads. The pneumatic agitation was very convenient in this case.List of Symbols D h–1 Dilution rate for continuous culture - h Residence time for continuous culture - dM/dt kg/(m3 · h) Rate of degradation of malic acid - dS/dt kg/(m3 · h) Rate of consumption of glucose - max h–1 Maximal specific rate of growth  相似文献   

10.
The effect of lead on the filtration rate of the zebra musselDreissena polymorpha was investigated, together with the accumulation of Pb in the soft tissues of the mussels. The NOEC-filtration was 116 g.l–1 (0,56 mol.l–1) and the EC50-filtration was 370 g.l–1 (1.79 mol.l–1). The NOEC-accumulation was the concentration found in the control water (1.4g.l–1). These experiments show that the EC50-filtration for Pb is similar to that for Cd, higher than that for Cu and lower than that for Zn. The water quality criteria for lead allow 25 g Pb.l–1 in surface water. This will not cause short-term effects. Long-term effects may, however, occur, since an accumulation of Pb as low as 16 g.l–1 was recorded in this study.  相似文献   

11.
Biochemical and biophysical parameters, including D1-protein turnover, chlorophyll fluorescence, oxygen evolution activity and zeaxanthin formation were measured in the marine seagrassZostera capricorni (Aschers) in response to limiting (100 mol·m–2·–1), saturating (350 mol·m–2·s–1) or photoinhibitory (1100 mol·m–2·s–1) irradiances. Synthesis of D1 was maximal at 350 mol·m–2·s–1 which was also the irradiance at which the rate of photosynthetic O2 evolution was maximal. Degradation of D1 was saturated at 350 mol·m–2·s–1. The rate of D1 synthesis at 1100 mol·m–2·s–1 was very similar to that at 350 mol·m–2·s–1 for the first 90 min but then declined. At limiting or saturating irradiance little change was observed in the ratio of variable to maximal fluorescence (Fv/Fm) measured after dark adaptation of the leaves, while significant photoinhibition occurred at 1100 mol·m–2·s–1. The proportion of zeaxanthin in the total xanthophyll pool increased with increasing irradiance, indicative of the presence of a photoprotective xanthophyll cycle in this seagrass. These results are consistent with a high level of regulatory D1 turnover inZostera under non-photoinhibitory irradiance conditions, as has been found previously for terrestrial plants.We would like to thank Professor Peter Böger (Department of Plant Biochemistry, University of Konstanz, Germany) for the kind gift of D1 antibodies. This work was partly supported by a University of Queensland Enabling Grant to CC.  相似文献   

12.
The CO2 gas exchange rates of the Central European perennial understory plantAsarum europaeum L. were measured in late autumn (October 30 to November 30) in its natural habitat day and night.During these measurements the temperature ranged from 0 to 15°C and the absolute air humidity from 3 to 10 mg H2O·1–1. Temperature and absolute air humidity over these ranges did not affect CO2 net assimilation which was determined almost entirely by quantum flux density.CO2 net assimilation was light saturated at about 100 M·m–2·s–1 quantum flux density. The uptake rate at this point was 4.3 mg·dm–2·h–1. The compensation point occurred at approximately 1 M·m–2·s–1.  相似文献   

13.
The dorsal skin of the leech Hirudo medicinalis was used for electrophysiological measurements performed in Ussing chambers. The leech skin is a tight epithelium (transepithelial resistance = 10.5±0.5 k· cm-2) with an initial short-circuit current of 29.0±2.9 A·cm-2. Removal of Na+ from the apical bath medium reduced short-circuit current about 55%. Ouabain (50mol·l-1) added to the basolateral solution, depressed the short-circuit current completely. The Na+ current saturated at a concentration of 90 mmol Na+·l-1 in the apical solution (K M=11.2±1.8 mmol·l-1). Amiloride (100 mol·l-1) on the apical side inhibited ca. 40% of the Na+ current and indicated the presence of Na+ channels. The dependence of Na+ current on the amiloride concentration followed Michaclis-Menten kinetics (K i=2.9±0.4 mol·l-1). The amiloride analogue benzamil had a higher affinity to the Na+ channel (K i=0.7±0.2 mol·l-1). Thus, Na+ channels in leech integument are less sensitive to amiloride than channels known from vertebrate epithelia. With 20 mmol Na+·l-1 in the mucosal solution the tissue showed an optimum amiloride-inhibitable current, and the amiloride-sensitive current under this condition was 86.8±2.3% of total short-circuit current. Higher Na+ concentrations lead to a decrease in amiloride-blockade short-circuit current. Sitmulation of the tissue with cyclic adenosine monophosphate (100 mol·l-1) and isobutylmethylxanthine (1 mmol·l-1) nearly doubled short-circuit current and increased amiloride-sensitive Na+ currents by 50%. By current fluctuation analysis we estimated single Na+ channel current (2.7±0.9 pA) and Na+ channel density (3.6±0.6 channels·m-2) under control conditions. After cyclic adenosine monophosphate stimulation Na+ channel density increased to 5.4±1.1 channels·m-2, whereas single Na+ channel current showed no significant change (1.9±0.2 pA). These data present a detailed investigation of an invertebrate epithelial Na+ channel, and show the similarities and differences to vertebrate Na+ channels. Whereas the channel properties are different from the classical vertebrate Na+ channel, the regulation by cyclic adenosine monophosphate seems similar. Stimulation of Na+ uptake by cyclic adenosine monophosphate is mediated by an increasing number of Na+ channels.Abbreviations slope of the background noise component - ADH antidiuretic hormone - cAMP cyclic adenosine monophosphate - f frequency - f c coner frequency of the Lorentzian noise component - Hepes N-hydroxyethylpiperazine-N-ethanesulphonic acid - BMX isobutyl-methylxanthine - i Na single Na+ channel current - I Na max, maximal inhibitable Na+ current - I SC short circuit current - K i half maximal blocker concentration - K M Michaelis constandard error of the mean - S (f) power density of the Lorentzian noise component - S 0 plateau value of the Lorentzian noise component - TMA tetramethylammonium - Trizma TRIS-hydroxymethyl-amino-methane - V max maximal reaction velocity - V T transepithelial potential - K half maximal blocker concentration  相似文献   

14.
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   

15.
The focus of this microcosm study was to monitor the performances of 17 butane-utilizing microcosms during a long-term (100–250 days) aerobic cometabolic depletion of chloroform (CF). The depletion of the contaminant began after a lag-time variable between 0 and 23 days. All microcosms quickly reached a pseudo steady-state condition, in terms of biomass concentration (with an average of 9.3 × 106 CFU ml–1), chloroform depletion rate (5 mol l–1 d–1) and butane utilization rate (730 mol l–1 d–1). After about 100 days of CF depletion, a sudden 5- to 7-fold increase of the chloroform rate was observed in two microcosms, where the highest amount of contaminant had been depleted. In one of these high-performing microcosms, an experiment of chloroform depletion in the absence of butane resulted in the depletion of a surprisingly high amount of contaminant (765 molCF kgdry soil–1 in 2 months) and in a marked selection of a single bacterial strain. Bioaugmentation assays conducted with the biomass selected in this microcosm and with a pure culture of the selected strain immediately resulted in very high chloroform depletion rates. Preliminary results of a study conducted with resting cells of the selected strain indicated that it can degrade chloroform concentrations up to 119 M (14.2 mg l–1) without any sign of substrate toxicity, and that it is able to transform vinyl chloride and 1,1,2-trichloroethane.  相似文献   

16.
Chloroplasts with high rates of photosynthetic O2 evolution (up to 120 mol O2· (mg Chl)-1·h-1 compared with 130 mol O2· (mg Chl)-1·h-1 of whole cells) were isolated from Chlamydomonas reinhardtii cells grown in high and low CO2 concentrations using autolysine-digitonin treatment. At 25° C and pH=7.8, no O2 uptake could be observed in the dark by high- and low-CO2 adapted chloroplasts. Light saturation of photosynthetic net oxygen evolution was reached at 800 mol photons·m-2·s-1 for high- and low-CO2 adapted chloroplasts, a value which was almost identical to that observed for whole cells. Dissolved inorganic carbon (DIC) saturation of photosynthesis was reached between 200–300 M for low-CO2 adapted chloroplasts, whereas high-CO2 adapted chloroplasts were not saturated even at 700 M DIC. The concentrations of DIC required to reach half-saturated rates of net O2 evolution (Km(DIC)) was 31.1 and 156 M DIC for low- and high-CO2 adapted chloroplasts, respectively. These results demonstrate that the CO2 concentration provided during growth influenced the photosynthetic characteristics at the whole cell as well as at the chloroplast level.Abbreviations Chl chlorophyll - DIC dissolved inorganic carbon - Km(DIC) coneentration of dissolved inorganic carbon required for the rate of half maximal net O2 evolution - PFR photon fluence rate - SPGM silicasol-PVP-gradient medium  相似文献   

17.
The light-dependent rate of photosystem-II (PSII) damage and repair was measured in photoautotrophic cultures of Dunaliella salina Teod. grown at different irradiances in the range 50–3000 mol photons · m–2· s–1. Rates of cell growth increased in the range of 50–800 mol photons·m–2·s–1, remained constant at a maximum in the range of 800–1,500 mol photons·m–2 ·s–1, and declined due to photoinhibition in the range of 1500–3000 mol photons·m–2·s–1. Western blot analyses, upon addition of lincomycin to the cultures, revealed first-order kinetics for the loss of the PSII reaction-center protein (D1) from the 32-kDa position, occurring as a result of photodamage. The rate constant of this 32-kDa protein loss was a linear function of cell growth irradiance. In the presence of lincomycin, loss of the other PSII reaction-center protein (D2) from the 34-kDa position was also observed, occurring with kinetics similar to those of the 32-kDa form of D1. Increasing rates of photodamage as a function of irradiance were accompanied by an increase in the steady-state level of a higher-molecular-weight protein complex ( 160-kDa) that cross-reacted with D1 antibodies. The steady-state level of the 160-kDa complex in thylakoids was also a linear function of cell growth irradiance. These observations suggest that photodamage to D1 converts stoichiometric amounts of D1 and D2 (i.e., the D1/D2 heterodimer) into a 160-kDa complex. This complex may help to stabilize the reaction-center proteins until degradation and replacement of D1 can occur. The results indicated an intrinsic half-time of about 60 min for the repair of individual PSII units, supporting the idea that degradation of D1 after photodamage is the rate-limiting step in the PSII repair process.Abbreviations Chl chlorophyll - PSI photosystem I - PSII photosystem II - D1 the 32-kDa reaction-center protein of PSII, encoded by the chloroplast psbA gene - D2 the 34-kDa reactioncenter protein of PSII, encoded by the chloroplast psbD gene - QA primary electron-accepting plastoquinone of PSII The work was supported by grant 94-37100-7529 from the US Department of Agriculture, National Research Initiative Competitive Grants Program.  相似文献   

18.
Summary Cellobiose oxidase from Phanerochaete chrysosporium was used for continuous monitoring of cellulase action on microcrystalline cellulose (Avicel). Two protocols are described, the parameter monitored being either the decline in electrode potential as ferricyanide is reduced or consumption of dioxygen. Most experiments used a commercial cellulase preparation from Trichoderma reesei and ferricyanide as acceptor. Within 1 min of an addition of cellulase, ferricyanide reduction reached a steady rate. This was converted into a rate of production of substrate for celobiose oxidase, in mol·min–1. Experiments were conducted either with a constant concentration of cellulase and increasing Avicel, or with constant Avicel and increasing cellulase. Kinetic analysis of the experiments with constant cellulase indicated a K mof 4.8 ± 1.0 (g cellulose)·1–1, which was close to the value predicted from binding studies. The specific activity of the cellulase was measured as 375±25 mol·(g cellulase)–1·min–1 in experiments with a high cellulose concentration, but was less than half this value when the cellulose was saturated with cellulase. The maximal rate of cellulose degradation was 9.6±1.3 mol·(g cellulose)–1·min–1.  相似文献   

19.
Lower concentrations of CuSO4 (25–75 M) in the MS medium supplemented with 0.1 mg l–1 IAA+5.0 mg l–1 Kn+500 mg l–1 CH+10 mg l–1 Cyst hyd enhanced the growth of regenerants of Dioscorea bulbifera L. CuSO4 (75 M) induced an appreciable diosgenin yield in the regenerants compared to those obtained on media without Cu. The presence of Cu thus seems to stimulate diosgenin production. The regenerants also differentiated bulbils on lower concentrations of Cu. At CuSO4 (100 M), however, cultures showed poor growth as well as a low diosgenin yield. Increased proline and protein contents were recorded in cultures grown on Cu-enriched media.  相似文献   

20.
Michael Luwe  Ulrich Heber 《Planta》1995,197(3):448-455
Spinach (Spinacia oleracea L.), broad bean (Vicia faba L.) and beech (Fagus sylvatica L.) plants were exposed to ozone at concentrations often measured in air during the summer months (120–300 g·m–3) and antioxidants were determined in the leaf tissue and in the aqueous phase of the cell wall, the apoplasm. Concentrations of both reduced ascorbate (AA) and its oxidized form, dehydroascorbate (DHA), showed the tendency to increase transiently in the apoplasm of spinach leaves 6–24 h after starting fumigation with ozone. In beech leaves, apoplasmic AA and DHA increased 3–7 d after beginning of treatment. At the very high concentration of 1600 g O3·m–3, an increase of apoplasmic AA was already measured after 1 d in beech leaves. Apparently, spinach and beech leaves respond to oxidative stress by increasing AA transport into the apoplasm and by accelerating DHA export. In contrast to these observations, DHA accumulated during 3 d of fumigation with only 120 g O3·m–3 in the apoplasm of broad bean leaves, while AA contents did not increase. After termination of fumigation, the extracellular redox state of ascorbate normalized within 1 d. Glutathione could not be detected in the apoplasm of any of the three leaf species. Intracellular AA changed its redox state in response to exposure to elevated concentrations of ozone. After 4–6 weeks of fumigation with 200–300 g O3·m–3 an increase of intracellular DHA was measured in beech leaves. At the same time, chlorophyll contents decreased and characteristic symptoms of ozone damage could be observed. However, no significant change in the redox state of apoplasmic ascorbate could be detected in beech leaves. Evidently, detoxification of ozone by apoplasmic AA was insufficient to protect the leaf tissue. Fumigation with a high ozone concentration (1600 g·m–3) caused an appreciable increase in the cellular contents of the oxidized forms of ascorbate and glutathione in beech leaves. Whereas in spinach leaves intracellular antioxidant contents and redox states were not altered during fumigation with 120–240 g O3·m–3, in broad bean leaves the intracellular DHA concentration increased and intracellular ascorbate became more oxidized after fumigation of the plants with 120 g O3·m–3. Apparently, broad bean leaves are more sensitive to ozone than beech and spinach leaves.Abbreviations AA ascorbate, reduced form - DHA ascorbate, oxidized form (dehydroascorbate) - FW fresh weight - GSH glutathione, reduced form - GSSG glutathione, oxidized form - IWF intercellular washing fluid - Vair intercellular air space volume of leaves - Vapo apoplasmic water volume of leaves This work was supported within the Sonderforschungsbereich 251 of the University of Würzburg.  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-23

京公网安备 11010802026262号