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 671 毫秒
1.
A general argument is made for the photochemical origins of life. A constant flux of free energy is required to maintain the organized state of matter called life. Solar photons are the unique source of the large amounts of energy probably require to initiate this organization and certainly required for the evolution of life to occur. The completion of this argument will require the experimental determination of suitable photochemical reactions. Our work shows that biogenetic porphyrins readily photooxidize substrates and emit hydrogen in the presence of a catalyst. These results are consistent with the Granick hypothesis, which relates a biosynthetic pathway to its evolutionary origin. We have shown that photoexcitation of ferrous ion at neutral pH with near ultraviolet light produces hydrogen with high quantum yield. This same simple system may reduce carbon dioxide to formaldehyde and further products. These reactions offer a solution to the dilemma confronting the Oparin-Urey-Miller model of the chemical origin of life. If carbon dioxide is the main form of carbon on the primitive earth, the ferrous photoreaction may provide the reduced carbon necessary for the formation of amino acids and other biogenic molecules. These results suggest that this progenitor of modern photosynthesis may have contributed to the chemical origins of life.  相似文献   

2.
Coccolithophorids are one of the dominant groups of marine phytoplankton. They are found in large numbers throughout the surface euphotic zone of the ocean, and are able to form large-scale blooms that persist for long periods of time. Coccolithophorid cells are covered by species-specific calcium carbonate crystals of various structures. In the process of calcification in coccolithophorids, Ca2+ is absorbed into cells from the culture medium, and a coccolith unit is formed inside the cell. Then, the coccolith unit extrudes to the cell surface where it is constructed into crystal layers. The formation of these crystals is regulated by cellular metabolism under different environmental conditions. The carbon biogeochemical cycle in the coccolithophorids involves both photosynthetic and calcification processes, which not only play an important role in population dynamics, but also in the global carbon cycle and climate change. However, one important question remains, namely, whether the relationship between photosynthesis and calcification is species-dependent. Previous studies have yielded controversial results, even in the same species. In this paper, we selected Pleurochrysis carterae, a coccolithophore species that frequently blooms in coastal areas, to study the relationship between calcification and photosynthesis. First, we studied population growth in a batch culture over several days. For batch cultures, P. carterae was inoculated into a 10 L bioreactor at an initial cell density of approximately 5 × 104 cells mL-1. The culture conditions were optimal for cell growth. Dissolved oxygen (DO) was detected during all the culture period, and the rate of photosynthetic oxygen evolution was calculated according the DO changes during the 12-h illumination period. Algal samples (10 mL) were collected during the population growth phases. The calcium carbonate content on the cell surface was determined each day by chemical titration. Next, we studied the relationship between photosynthesis and calcification at the cellular level by observing patterns of recalcification during a 12-h period. In this study, non-calcified cells were obtained by decalcifying calcified cells collected during the exponential growth period in MES-NaOH buffer solution (pH 5.5). The non-calcified cells were inoculated into culture media containing different concentrations of Ca2+ (0, 5, 20, 40, 50, or 100 mg L-1). The rate of recalcification was determined by microscopic analyses in which the number of recalcified cells per 100 cells was counted at 0, 3, 6, 9, and 12 h of culture. Ca2+ absorbed into the cell was detected by measuring the fluorescence intensity of Fluo-3/AM labeled Ca2+. The rate of photosynthetic oxygen evolution in the non-calcified cell cultures was detected by measuring the changes in dissolved oxygen during the 12-h illumination period. The results showed that during the population growth period, the rate of photosynthetic oxygen evolution was inversely related to the calcium carbonate content per cell. When the amount of calcium carbonate on the cell surface increased, the relative photosynthetic ability (the rate of photosynthetic oxygen evolution) decreased, and vice versa. Both recalcification rates and photosynthetic oxygen evolution were affected by the extracellular calcium concentration. Non-calcified cells showed different recalcification abilities at different extracellular Ca2+ concentrations. The recalcification rate of non-calcified cells was positively correlated with the extracellular calcium concentration when [Ca2+] in the medium ranged from 0 to 100 mg L-1. However, photosynthetic oxygen evolution was suppressed at higher cell calcification rates, especially when extracellular [Ca2+] was 50–100 mg L-1. Our analyses of the population growth process and the cell recalcification process confirmed that photosynthesis is inversely related to calcification in P. carterae.  相似文献   

3.
Anacystis nidulans grown under high and low light, 100 and 10 μE m?2 s?1, respectively, was analyzed with respect to chlorophyll/P700, phycobiliproteins/P700, chlorophyll/cell, and oxygen evolution parameters. The photosynthetic unit sizes of this cyanobacterium, measured as the ratio of total chromophores (chlorophyll and bilin) to P700, were shown to be similar to those of higher plants and green algae. High light grown cells possessed a photosynthetic unit consisting of a core of 157 ± 6 chlorophyll a molecules per P700 associated with a light harvesting system of 95 ± 3.5 biliprotein chromophores. Low light grown cells had substantially more biliprotein chromophores per P700 (125 ± 3.1) than high light cells, but showed no significant difference in the numbers of chlorophyll a molecules per P700 (149 ± 4). Analyses of aqueous biliprotein extracts indicate that low light grown cells produce proportionately more phycocyanin relative to allophycocyanin than high light cells. Calculations of the molecular weight of biliproteins per P700 suggest that there is less than one phycobilisome per reaction center I under both growth conditions. Differences in chlorophyll/cell ratios and oxygen evolution characteristics were also observed. High light cells contain 6.3 × 10?12 mg chlorophyll cell?1, while low light grown cells contain 12.8 × 10?12 mg chlorophyll cell?1. Photosynthetic oxygen evolution rate vs. light intensity curves indicate that high light grown cells reach maximal levels of oxygen evolution at higher light intensity than low light grown cells. Maximal rates of oxygen evolution were 16.6 μmol oxygen min?1 (mg chlorophyll)?1 for high and 8.4 μmol oxygen min?1 (mg chlorophyll)?1 for low light cells. Maximal oxygen evolution rates per cell were equivalent for both cell types, although the amount of P700 per cell was lower in high light cells. High light grown cells are therefore capable of producing more oxygen per reaction center I than low light grown cells.  相似文献   

4.
Anatomical and physiological leaf characteristics and biomass production of Fatsia japonica plants were studied. Plants were grown in a growth chamber at 300 μmol m-2 s-1 (high light) and 50 μmol m-2 s-1 (low light) photosynthetic photon flux density. Plants grown under high light showed a net maximum photosynthetic rate 44% higher than plants grown under low light; the light compensation point and the light saturation point were also higher in high-light plants. Photosynthetic oxygen evolution in isolated chloroplasts was about 40% higher in high-light plants. However, chlorophyll content on a dry weight basis, on a leaf area basis, and per chloroplast was greater in plants grown under low light. Leaf thickness in high-light plants was 13% higher than in low-light plants. The number of chloroplasts was 30% higher in high-light leaves, while chloroplast size was only slightly higher. Chloroplast ultrastructure was also affected by light. Leaf dry weight, leaf area, and biomass production per plant were drastically reduced under low light. Thus, F. japonica is a plant that is able to acclimate to different photosynthetic photon flux density by altering its anatomical and physiological characteristics. However, low-light acclimation of this plant has a considerable limiting effect on biomass production.  相似文献   

5.
Ozone Inhibition of Photosynthesis in Chlorella sorokiniana   总被引:2,自引:0,他引:2       下载免费PDF全文
Exposure of Chlorella sorokiniana (07-11-05) to ozone inhibits photosynthesis. In this study, the effects of ozone on O2 evolution and fluorescence yields are used to characterize this inhibition. At an ozone dose of about 3 micromoles delivered to 2 × 109 cells, the photosynthetic rate of the cells is inhibited 50%, as indicated by a decrease in bicarbonate-stimulated O2 evolution (control rate, 1.4 ± 0.3 × 10−15 moles per cell per minute).  相似文献   

6.
A growth regulator (G; 4-ethyl-1-hydroxy-4,8,8,10,10 pentamethyl-7,9-dioxo-2,3 dioxyabicyclo (4.4.0) decene-5) from Eucalyptus grandis (Maiden) reduced stomatal conductance and also photosynthetic capacity when fed through the transpiration stream of detached leaves. The concentration of G required for this effect was high (10−4 molar), but the amount of G taken up (dose) was below the level which has previously been found in E. grandis leaves. Similar effects were observed in detached leaves of Xanthium strumarium L. though almost 10 times more G was required. G reduced CO2-dependent O2 evolution from isolated cells of X. strumarium. In spinach (Spinacia oleracea L.) chloroplasts, electron transport through photosystem II was reduced by G. It is proposed that G affects stomatal conductance and photosynthesis by reducing photosystem II activity in both the guard cell chloroplasts and mesophyll cell chloroplasts.  相似文献   

7.
CO2 exchange rates (CO2 evolution) of late-maturing cv. Cal Red peaches, exposed to different photon flux densities, were simulated from 24 days after flowering (DAF) until harvest by using light and temperature response curves measured on attached fruits in the field at biweekly intervals. The daily patterns of dark respiration rates per unit dry weight indicated their dependence on temperatures. Fruit CO2 exchange rates in light were also affected by photosynthetic photon flux densities. Daily photosynthetic rates per unit dry weight and per fruit were significantly lower in shaded fruits receiving 7% of the full daily sunlight compared to fruits exposed to 35% sunlight. However, the difference in photosynthetic rates in peach fruits receiving 21 and 35% of total daily sunlight was small. Within the last 4 weeks before harvest, weekly carbohydrate requirements for the production of dry matter rose rapidly in cv. Cal Red peaches and were related to high carbohydrate accumulations, especially of sucrose, in the peach mesocarp. Weekly photosynthetic contribution of late-maturing cv. Cal Red peaches to these carbohydrate accumulations increased up to 115 DAF. A decline in photosynthetic contributions between 115 DAF and harvest was related to decreasing photosynthetic activities in association with declining chlorophyll contents. Photosynthesis of late-maturing cv. Cal Red peaches provided 3–9% of the weekly fruit carbohydrate requirements early in the season and 8–15% in the midseason depending on fruit exposure to light. Photosynthesis of mature fruits contributed 3–5% of the total fruit carbohydrate requirements. Since fruit photosynthetic rates approach saturation at a photosynthetic photon flux density of about 600 μmol m2 s?2, the difference in weekly photosynthetic contributions was small between exposed and partially exposed (35 and 21% sunlight, respectively) peach fruits. However, a shaded fruit (7% sunlight) supplied significantly less of its weekly carbohydrate requirements through photosynthesis compared to exposed fruits. During the growing period of 24 DAF until harvest, dry matter accumulation of latematuring cv. Cal Red peaches accounted for 78% of the total carbohydrate requirements and 22% was used in respiration. Fruit photosynthesis of shaded peach fruit, partially exposed fruit and exposed fruit (receiving 7. 21 and 35% of full sunlight over the day, respectively) contributed 5. 8 and 9%, respectively, of the total growth and maintenance carbohydrate requirements during the growing season.  相似文献   

8.
CO2 exchange rates per unit dry weight, measured in the field on attached fruits of the late-maturing Cal Red peach cultivar, at 1200 μmol photons m?2S?1 and in dark, and photosynthetic rates, calculated by the difference between the rates of CO2 evolution in light and dark, declined over the growing season. Calculated photosynthetic rates per fruit increased over the season with increasing fruit dry matter, but declined in maturing fruits apparently coinciding with the loss of chlorophyll. Slight net fruit photosynthetic rates ranging from 0. 087 ± 0. 06 to 0. 003 ± 0. 05 nmol CO2 (g dry weight)?1 S?1 were measured in midseason under optimal temperature (15 and 20°C) and light (1200 μmol photons m?2 S?1) conditions. Calculated fruit photosynthetic rates per unit dry weight increased with increasing temperatures and photon flux densities during fruit development. Dark respiration rates per unit dry weight doubled within a temperature interval of 10°C; the mean seasonal O10 value was 2. 03 between 20 and 30°C. The highest photosynthetic rates were measured at 35°C throughout the growing season. Since dark respiration rates increased at high temperatures to a greater extent than CO2 exchange rates in light, fruit photosynthesis was apparently stimulated by high internal CO2 concentrations via CO2 refixation. At 15°C, fruit photosynthetic rates tended to be saturated at about 600 μmol photons m?2 S?1. Young peach fruits responded to increasing ambient CO2 concentrations with decreasing net CO2 exchange rates in light, but more mature fruits did not respond to increases in ambient CO2. Fruit CO2 exchange rates in the dark remained fairly constant, apparently uninfluenced by ambient CO2 concentrations during the entire growing season. Calculated fruit photosynthetic rates clearly revealed the difference in CO2 response of young and mature peach fruits. Photosynthetic rates of younger peach fruits apparently approached saturation at 370 μl CO21?2. In CO2 free air, fruit photosynthesis was dependent on CO2 refixation since CO2 uptake by the fruits from the external atmosphere was not possible. The difference in photosynthetic rates between fruits in CO2-free air and 370 μl CO2 1?1 indicated that young peach fruits were apparently able to take up CO2 from the external atmosphere. CO2 uptake by peach fruits contributed between 28 and 16% to the fruit photosynthetic rate early in the season, whereas photosynthesis in maturing fruits was supplied entirely by CO2 refixation.  相似文献   

9.
In bright sunlight photosynthetic activity is limited by the enzymatic machinery of carbon dioxide assimilation. This supererogation of energy can be easily visualized by the significant increases of photosynthetic activity under high CO2 conditions or other metabolic strategies which can increase the carbon flux from CO2 to metabolic pools. However, even under optimal CO2 conditions plants will provide much more NADPH + H+ and ATP that are required for the actual demand, yielding in a metabolic situation, in which no reducible NADP+ would be available. As a consequence, excited chlorophylls can activate oxygen to its singlet state or the photosynthetic electrons can be transferred to oxygen, producing highly active oxygen species such as the superoxide anion, hydroxyl radicals and hydrogen peroxide. All of them can initiate radical chain reactions which degrade proteins, pigments, lipids and nucleotides. Therefore, the plants have developed protection and repair mechanism to prevent photodamage and to maintain the physiological integrity of metabolic apparatus. The first protection wall is regulatory energy dissipation on the level of the photosynthetic primary reactions by the so-called non-photochemical quenching. This dissipative pathway is under the control of the proton gradient generated by the electron flow and the xanthophyll cycle. A second protection mechanism is the effective re-oxidation of the reduction equivalents by so-called “alternative electron cycling” which includes the water-water cycle, the photorespiration, the malate valve and the action of antioxidants. The third system of defence is the repair of damaged components. Therefore, plants do not suffer from energy shortage, but instead they have to invest in proteins and cellular components which protect the plants from potential damage by the supererogation of energy. Under this premise, our understanding and evaluation for certain energy dissipating processes such as non-photochemical quenching or photorespiration appear in a quite new perspective, especially when discussing strategies to improve the solar energy conversion into plant biomass.  相似文献   

10.
Acclimation to three photon flux densities (10, 35, 180 μE.m?2.s?1) was determined in laboratory cultures of Porphyridium purpureum Bory, Drew and Ross. Cultures grown at low, medium, and high PPFDs had compensation points of <3, 6, and 20 μE-m?2.s?1, respectively, and saturating irradiances in the initial log phase of 90, 115, 175 μE.m?2.s?1 and up to 240 μE.m?2.s?1 in late log phase. High light cells had the smallest photosynthetic unit size (phycobiliproteins plus chlorophyll), the highest photosynthetic capacity, and the highest growth rates. Photosystem I reaction centers (P700) per cell remained proportional to chlorophyll at ca. 110 chl / P700. However, phycobiliprotein content decreased as did the phycobilisome number (ca. 50%) in high light cells, where as the phycobilisome size remained the same as in medium and low light cells. We concluded that acclimation of this red alga to varied PPFDs was manifested by the plasticity of the photosystem II antennae with little, if any, effect noted on photosystem I.  相似文献   

11.
Air-grown cells of Porphyridium purpurem contain appreciable carbonic-anhydrase activity, comparable to that in air-grown Chlamydomonas reinhardtii, but activity is repressed in CO2-grown cells. Assay of carbonic-anhydrase activity in intact cells and cell extracts shows all activity to be intracellular in Porphyridium. Measurement of inorganic-carbon-dependent photosynthetic O2 evolution shows that sodium ions increase the affinity of Porphyridium cells for HCO 3 - . Acetazolamide and ethoxyzolamide were potent inhibitors of carbonic anhydrase in cell extracts but at pH 5.0 both acetazolamide and ethoxyzolamide had little effect upon the concentration of inorganic carbon required for the half-maximal rate of photosynthetic O2 evolution (K0.5[CO2]). At pH 8.0, where HCO 3 - is the predominant species of inorganic carbon, the K0.5 (CO2) was increased from 50 M to 950 M in the presence of ethoxyzolamide. It is concluded that in air-grown cells of Porphyridium. HCO 3 - is transported across the plasmalemma and intracellular carbonic anhydrase increases the steady-state flux of CO2 from inside the plasmalemma to ribulose-1,5-bisphosphate carboxylase-oxygenase by catalysing the interconversion of HCO 3 - and CO2 within the cell.Abbreviations AZ acetazolamide - EZ ethoxyzolamide - K0.5[CO2] half-maximal rate of photosynthetic O2 evolution  相似文献   

12.
Abstract A simple mechanical method for the rapid isolation of chloroplasts with high rates of photosynthesis from young leaves of oat (Avena sativa L.) was described. The photosynthetic activity of these chloroplasts was stable for at least 2 h with rates of CO2-dependent O2 evolution of 30–40 μmol g 1 Chl s 1. The photosynthetic properties of these chloroplasts were similar to those reported for spinach and pea chloroplasts isolated by mechanical disruption. The pH optimum for photosynthetic O2 evolution was pH 7.6. The induction time was 0.5–2 min. Maximal rates of photosynthetic O2 evolution in these chloroplast preparations were obtained in the absence of both divalent cations and EDTA. Addition of divilent cations strongly inhibited photosynthesis which could be partially restored by the subsequent addition of EDTA. But when these cations were not present in the assay medium the addition of EDTA greater than 1 mol m 3 decreased photosynthetic activity. The optimal orthophosphate concentration required for photosynthesis in these chloroplast preparations was 0.2–0.3 mol m 3. In contrast, the addition of pyrophosphate either in the light or dark inhibited photosynthesis. In a comparative study, chloroplasts were also isolated from oat and wheat (Triticum aestivum L., cultivar Hybrid C306) protoplasts. These chloroplast preparations were found to have properties similar to those determined for oat chloroplasts isolated by the mechanical method reported above.  相似文献   

13.
Abstract Photon requirements for growth (φg?1) of the diatom Phaeodactylum tricornutum were determined under nutrient-sufficient conditions at two photon flux densities corresponding to light limited and near-saturating conditions for growth. The value of φg?1 based on assimilated carbon was light-dependent and varied from 8.8 to 14.0 mol photon mol C?1 with the minimum value at the lowest photon flux density. These results are lower than might be predicted for microalgal growth based on the Z scheme of photosynthesis. Conversion of these values for carbon fixation to estimates based on oxygen evolution is problematical due to uncertainty over the appropriate assimilatory quotient (Qa= mol O2 mol C?1). Minimum values based on oxygen evolution rates ranged from 6.2 to 7.6 mol photon mol O2?1 using a Qa of 1.41 mol O2 mol C?1 obtained by Myers (1980). These estimates are similar to our previous measurements for photosynthesis and indicate a high efficiency for light energy transforming reactions during growth. The values of (φg?1 obtained in this work indicate a number of inadequacies in our understanding of the energetics of microalgal growth and are inconsistent with our present knowledge of photosynthetic energy coupling in plant cells.  相似文献   

14.
Several experiments were conducted to understand better the physiological mechanisms underlying growth inhibition of the dinoflagellate Gonyaulax polyedra Stein due to small-scale turbulence shear. To measure photosynthetic 14C uptake, a “phytoplankton wheel” device for rotating cultures in closed bottles was used. Turbulence was quantified biologically in the bottles by comparing growth inhibition with that in cultures with constant shear between a fixed cylinder and an outer concentric rotating cylinder (a stable Couette flow). At saturating irradiances, particulate photosynthesis (Psat) or photosynthesis per unit chlorophyll (PBsat) were not inhibited completely at the highest turbulence level (26.6 rad.s?1), and photosynthesis was less sensitive than growth. Photosynthesis per cell (PCsat) was increased by turbulence. In three experiments on the effects of turbulence on photosynthesis versus irradiance curves, the slope of the curve, α, for particulate photosynthesis at limiting irradiances did not change. Photosynthesis per unit chlorophyll per unit irradiance (αB) decreased at high (but not intermediate) turbulence levels. Photosynthesis per cell per unit irradiance, αC, increased with turbulence, suggesting an increase in photosynthetic efficiency in turbulent cultures. In two of the three experiments, respiration rates increased with turbulence, and in one experiment excretion of photosynthetically fixed 14C was not affected by motion. Ratios of accessory pigments to chlorophyll a did not change with turbulence, but pigments per cell and per dry weight increased with turbulence. These findings suggest little or no disruption of the photosynthetic apparatus. When turbulence was applied for 1 week, β-carotene increased while peridinin and diadinoxanthin decreased, suggesting inhibition of synthesis of these latter pigments by prolonged turbulence. Since cell numbers did not increase or decreased during turbulent 72–h incubations, cell division was inhibited and also the cells were very much enlarged. Increases in αC per cell suggest that, in the sea, photo synthetic metabolism can persist efficiently without cell division during turbulent episodes. After turbulence ceases or reaches low levels again, cells can then divide and blooms may form. Thus, blooms can come or go fairly rapidly in the ocean depending on the degree of wave- and wind-induced turbulence.  相似文献   

15.
Action Spectra for Guard Cell Rb Uptake and Stomatal Opening in Vivia faba   总被引:5,自引:5,他引:0  
Abaxial epidermal strips, containing guard cells as the only viable cells, were prepared from leaves of Vicia faba following a period in darkness, and floated, under CO2-free air, on 2 mm RbCl + 0.1 mm CaCl2 labeled with 86Rb+. Under white light (high pressure mercury vapor lamp), stomatal opening in these strips approached its maximum at less than 0.02 calorie per square centimeter per minute. Under light of different wavelengths, 20 nanometers apart, and at a low quantum flux density of 7 × 1014 quanta per square centimeter per second, Rb+ uptake and stomatal opening were activated only in the blue and long ultraviolet regions, with a peak at 420 to 460 nanometers. The action spectrum suggests that the underlying process is not photosynthesis. At higher quantum flux density (38 × 1014 quanta per square centimeter per second), uptake and opening also responded to red (600-680 nanometers) and somewhat to green light, with a minimum at 540 to 560 nanometers, indicating a possible involvement of the photosynthetic process. This light-induced opening appeared not to be mediated by a lowering of CO2 concentration, since CO2-free air was used in all treatments and controls. Stomatal opening paralleled Rb+ uptake in all cases. This constitutes further evidence for the potassium transport hypothesis of stomatal movement.  相似文献   

16.
Homocontinuous cultures of the cyanobacterium Anacystis nidulans (syn. Synechococcus sp. PCC 6301) were grown at white light intensities of 2 and 20 W/m2, and supplied with 0.03 and 3 % CO2 enriched air. The mutual influence of these growth factors on the development of the photosynthetic apparatus was studied by analyses of the pigment content, by low temperature absorbance and fluorescence spectroscopy, by analyses of oxygen evolution light-saturation curves, and by SDS PAGE of isolated phycobilisomes. The two growth factors, light and CO2, distinctly affect the absorption cross section of the photosynthetic apparatus, which is expressed by its pigment pattern, excitation energy distribution and capacity. In response to low CO2 concentrations, the phycocyanin / allophycocyanin ratios were lower and one linker polypeptide L30R, of the phycobilisomes was no longer detectable in SDS PAGE. Apparently, low CO2 adaptation results in shorter phycobilisome rods. Specifically, upon adaptation to low light intensities, the chlorophyll and the phycocyanin content on a per cell basis increase by about 50% suggesting a parallel increase in the amount of phycobilisomes and photosystem core-complexes. Low light adaptation and low CO2 adaptation both cause a shift of the excitation energy distribution in favor of photosystem I. Variations in the content of the “anchor” polypeptides L60CM and L75CM are possibly related to changes in the excitation energy transfer from phycobilisomes to the photosystem II and photosystem I core-complexes.  相似文献   

17.
Havaux M 《Plant physiology》1989,89(1):286-292
The photoacoustic technique was used to monitor thermal deexcitation of the photosynthetic pigments in intact pea leaves (Pisum sativum L.) submitted to photoinhibitory treatments. When the leaves were exposed to photon flux densities above 1000 micromoles per square meter per second, the amplitude of the photothermal component of the in vivo photoacoustic signal strongly increased. This high-light-induced stimulation of nonradiative energy dissipation (heat emission) was accompanied by an inverse change in the O2 evolution activity and in the steady state emission of 685 nanometer chlorophyll fluorescence. The time course of these effects was shown to be very rapid, with a t1/2 of around 15 minutes. When high-light-treated leaves were readapted to the dark, the heat emission changes were reversed, following somewhat slower kinetics. A reversible increase in the rate of light energy dissipation via radiationless transitions could be a photoprotective mechanism eliminating excess excitation energy from the photosynthetic reaction centers. Interestingly, this process does not operate at temperatures below about 12°C.  相似文献   

18.
Physiological evidence was obtained for a light shielding role for gas vacuoles inMicrocystis aeruginosa Kuetz. emend. Elenkin, by comparing photosynthetic oxygen evolution, growth behaviour and pigment composition of cells with intact or collapsed gas vacuoles. The oxygen evolution rates were strongly dependent on cell concentration, a maximum rate for cells with intact gas vacuoles occurring at about 1.4×109 cells/ml and for cells with collapsed gas vacuoles at about 2.5×109 cells/ml. By using light saturation curves for oxygen evolution, it was estimated that at low light intensities up to 30% of the photosynthetically useable light was shielded at a cell concentration of 6×108 cells/ml. Collapsing the gas vacuoles twice daily did not alter the initial growth rate of the cultures, but enabled them to reach a higher final cell density. Collapsing of gas vacuoles during growth for about four generations resulted in a lower level of all acetone soluble pigments with a greater relative reduction in carotenoids than in chlorophyll a. Collapse of the gas vacuoles does not alter the cell volume. Various optical interactions which could account for light shielding are discussed.  相似文献   

19.
Nitrate-limited chemostat cultures of Selenastrum minutum Naeg. Collins (Chlorophyta) were used to determine the effects of nitrogen addition on photosynthesis, dark respiration, and dark carbon fixation. Addition of NO3 or NH4+ induced a transient suppression of photosynthetic carbon fixation (70 and 40% respectively). Intracellular ribulose bisphosphate levels decreased during suppression and recovered in parallel with photosynthesis. Photosynthetic oxygen evolution was decreased by N-pulsing under saturating light (650 microeinsteins per square meter per second). Under subsaturating light intensities (<165 microeinsteins per square meter per second) NH4+ addition resulted in O2 consumption in the light which was alleviated by the presence of the tricarboxylic acid cycle inhibitor fluoroacetate. Addition of NO3 or NH4+ resulted in a large stimulation of dark respiration (67 and 129%, respectively) and dark carbon fixation (360 and 2080%, respectively). The duration of N-induced perturbations was dependent on the concentration of added N. Inhibition of glutamine 2-oxoglutarate aminotransferase by azaserine alleviated all these effects. It is proposed that suppression of photosynthetic carbon fixation in response to N pulsing was the result of a competition for metabolites between the Calvin cycle and nitrogen assimilation. Carbon skeletons required for nitrogen assimilation would be derived from tricarboxylic acid cycle intermediates. To maintain tricarboxylic acid cycle activity triose phosphates would be exported from the chloroplast. This would decrease the rate of ribulose bisphosphate regeneration and consequently decrease net photosynthetic carbon accumulation. Stoichiometric calculations indicate that the Calvin cycle is one source of triose phosphates for N assimilation; however, during transient N resupply the major demand for triose phosphates must be met by starch or sucrose breakdown. The effects of N-pulsing on O2 evolution, dark respiration, and dark C-fixation are shown to be consistent with this model.  相似文献   

20.
Relations between irradiance (I) and lichen growth were investigated for five macro‐lichens growing at two sites in Sweden. The lichens represented different mycobiont–photobiont associations, two morphologies (foliose, fruticose) and two life forms (epiphytic, terricolous). The lichens were transplanted at two geographically distant sites in Sweden (1000 km apart) from Sept 1995 to Sept 1996 in their typical microhabitats, where microclimate and growth were followed. Between April/May and Sept 96, the terricolous species had a dry matter gain of 0·2 to 0·4 g (g DW)–1 and the epiphytes 0·01 to 0·02 g (g DW)–1. When related to area, growth amounted to 30 to 70 g m?2 for the terricolous species and to 1 to 4 g m?2 for the epiphytes. There was a strong correlation between growth and intercepted irradiance when the lichens were wet (Iwet), with 0·2 to 1·1 g lichen dry matter being produced per MJ solar energy. Across the 10 sets of transplants, light use efficiencies of dry matter yield (e) ranged between 0·5 and 2%, using an energy equivalent of 17·5 kJ g?1 of lichen dry matter. The higher productivity of the terricolous species was due to longer periods with thallus water contents sufficient for metabolic activity and because of the higher mean photon flux densities of their microhabitat. A four‐fold difference in photosynthetic capacity among the species was also important. It is concluded that lichen dry matter gain was primarily related to net carbon gain during metabolically active periods, which was determined by light duration, photon flux density and photosynthetic capacity.  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