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 15 毫秒
1.
Summary The vapor pressure difference between H2 18O and H2 16O is the reason for the accumulation of the heavy molecule in transpiring leaves. Since photosynthesis on land is the main source of atmospheric oxygen, this mechanism is important for the remarkable enrichment of18O in atmospheric O2 (Dole effect). Using a simple box model for transpiring leaves a quantitative understanding of the isotope fractionation is possible which is well confirmed by the results of model experiments as well as by measurements on trees. Maximum enrichment of H2 18O in the water of leaves (relative to soil water) is 25 (theoretically, for dry air) and was found under natural conditions to be 21 (for 28 % relative humidity); minimum theoretical enrichment is zero (observed 2.5 ).  相似文献   

2.
Measurements were made of the concentration and stable oxygen isotopic ratio of carbon dioxide in air samples collected on a diurnal basis at two heights within a Pinus resinosa canopy. Large changes in CO2 concentration and isotopic composition were observed during diurnal time courses on all three symple dates. In addition, there was strong vertical stratification in the forest canopy, with higher CO2 concentrations and more negative 18O values observed closer to the soil surface. The observed daily increases in 18O values of forest CO2 were dependent on relative humidity consistent with the modelled predictions of isotopic fractionation during photosynthetic gas exchange. During photosynthetic gas exchange, a portion of the CO2 that enters the leaf and equilibrates with leaf water is not fixed and diffuses back out of the leaf with an altered oxygen isotopic ratio. The oxygen isotope ratio of CO2 diffusing out of a leaf depends primarily on the 18O content of leaf water which changes in response to relative humidity. In contrast, soil respiration caused a decline in the 18O values of forest CO2 at night, because CO2 released from the soil has equilibrated with soil water which has a lower 18O content than leaf water. The observed relationship between diurnal changes in CO2 concentration and oxygen isotopic composition in the forest environment were consistent with a gas mixing model that considered the relative magnitudes of CO2 fluxes associated with photosynthesis, respiration and turbulent exchange between the forest and the bulk atmosphere.  相似文献   

3.
ATMOSPHERIC oxygen is not in equilibrium with sea water with respect to the isotope exchange illustration but has an 18O excess of about 22‰ compared to sea water1. This could be due to isotope fractionation during respiration2. Another large contribution to the effect has been overlooked up to now. Photosynthesis on land takes place in transpiring leaves, where the difference in the vapour pressure of 16OH2 and 18OH2 concentrates the heavy molecules in their stationary water content. Since the free oxygen stems from the water in which photosynthesis takes place3–8 (with only a very small shift in isotopic composition9), photosynthesis on land is an 18O source for atmospheric O2. We have begun to study this effect quantitatively.
  相似文献   

4.
The natural abundance hydrogen-isotope composition of leaf water ( ) and leaf organic matter ( D org ) was measured in leaves of C3 and C4 dicotyledons and monocotyledons. The value of leaf water showed a marked diurnal variation, greatest enrichment being observed about midday. However, this variation was greater in the more slowly transpiring C4 plants than in C3 plants under comparable environmental conditions. A model based on analogies with a constant feed pan of evaporating water was developed and the difference between C3 and C4 plants expressed in terms of either differences in kinetic enrichment or different leaf morphology. Microclimatic and morphological features of the leaves which may be associated with this factor are discussed. There was no daily excursion in the D org value in leaves of either C3 or C4 plants. When D org values were referenced to the mean values during the period of active photosynthesis, the discrimination against deuterium during photosynthetic metabolism (D) was greater in C3 plants (-117 to -121) than in C4 plants (-86 to -109).These results show that the different water use strategies of C3 and C4 plants are responsible for the measured difference in deuterium-isotope composition of leaf water. However, it is unlikely that these physical processes account fully for the differences in hydrogen-isotope composition of the products of C3 and C4 photosynthetic metabolism.Symbols Hydrogen-isotope composition of leaf water - D org hydrogen-isotope composition of leaf organic matter  相似文献   

5.
In order to assess the ability of Porites corals to accurately record environmental variations, high-resolution (weekly/biweekly) coral 18O records were obtained from four coral colonies from the northern Gulf of Aqaba, which grew at depths of 7, 19, 29, and 42 m along one transect. Adjacent to each colony, hourly temperatures, biweekly salinities, and monthly 18O of seawater were continuously recorded over a period of 14 months (April 1999 to June 2000). Contrary to water temperature, which shows a regular and strong seasonal variation and change with depth, seawater 18O exhibits a weak seasonality and little change with depth. Positive correlations between seawater 18O and salinity were observed. The two parameters were related to each other by the equation 18O Seawater (, VSMOW) = 0.281 × Salinity – 9.14. The high-resolution coral 18O records from this study show a regular pattern of seasonality and are able to capture fine details of the weekly average temperature records. They resolve more than 95% of the weekly average temperature range. On the other hand, attenuation and amplification of coral seasonal amplitudes were recorded in deep, slow-growing corals, which were not related to environmental effects (temperature and/or seawater 18O) or sampling resolution. We propose that these result from a combined effect of subannual variations in extension rate and variable rates of spine thickening of skeletal structures within the tissue layer. However, no smoothing or distortion of the isotopic signals was observed due to calcification within the tissue layer in shallow-water, fast-growing corals. The calculations from coral 18O calibrations against the in situ measurements show that temperature (T) is related to coral 18O ( c ) and seawater 18O ( w ) by the equation T (°C) = –5.38 ( c w ) –1.08. Our results demonstrate that coral 18O from the northern Gulf of Aqaba is a reliable recorder of temperature variations, and that there is a minor contribution of seawater 18O to this proxy, which could be ignored.  相似文献   

6.
We analysed the stable isotope composition of emitted N2O in a one-year field experiment (June 1998 to April 1999) in unfertilized controls, and after adding nitrogen by applying slurry or mineral N (calcium ammonium nitrate). Emitted N2O was analysed every 2–4 weeks, with additional daily sampling for 10 days after each fertilizer application. In supplementary soil incubations, the isotopic composition of N2O was measured under defined conditions, favouring either denitrification or nitrification. Soil incubated for 48 h under conditions favouring nitrification emitted very little N2O (0.024 mol gdw –1) and still produced N2O from denitrification. Under denitrifying incubation conditions, much more N2O was formed (0.91 mol gdw –1 after 48 h). The isotope ratios of N2O emitted from denitrification stabilized at 15N = –40.8 ± 5.7 and 18O = 2.7 ± 6.3. In the field experiment, the N2O isotope data showed no clear seasonal trends or treatment effects. Annual means weighted by time and emission rate were 15N = –8.6 and 18O = 34.7 after slurry application, 15N = –4.6 and 18O = 24.0 after mineral fertilizer application and 15N = –6.4 and 18O = 35.6 in the control plots, respectively. So, in all treatments the emitted N2O was 15N-depleted compared to ambient air N2O (15N = 11.4 ± 11.6, 18O = 36.9 ± 10.7). Isotope analyses of the emitted N2O under field conditions per se allowed no unequivocal identification of the main N2O producing process. However, additional data on soil conditions and from laboratory experiments point to denitrification as the predominant N2O source. We concluded (1) that the isotope ratios of N2O emitted from the field soil were not only influenced by the source processes, but also by microbial reduction of N2O to N2 and (2) that N2O emission rates had to exceed 3.4 mol N2O m–2 h–1 to obtain reliable N2O isotope data.  相似文献   

7.
We determined the 18O and 13C composition of the same fixed growth increment in severalPorites lutea coral skeletons from Phuket, South Thailand. Skeletal growth rate and 18O are inversely related. We explain this in terms of McConnaughey's kinetic isotopic disequilibria model. Annual trends in 18O cannot be solely explained by observed variations in seawater temperature or salinity and may also reflect seasonal variations in calcification rate. Coral tissue chlorophylla content and 13C of the underlying 1 mm of skeleton are positively related, suggesting that algal modification of the dissolved inorganic carbonate pool is the main control on skeletal 13C. However, in corals that bleached during a period of exceptionally high seawater temperatures in the summer of 1991, 13C of the outer 1 mm of skeleton and skeletal growth rate (over 9 months up to and including the bleaching event) are inversely related. Seasonal variations in °13C may reflect variations in calcification rate, zooxanthellae photosynthesis or in seawater 13C composition. Bleached corals had reduced calcification over the 9-month period up to and including the bleaching event and over the event they deposited carbonate enriched in13C and18O compared with unaffected corals. However, calcification during the event was limited and insufficient material was deposited to influence significantly the isotopic signature of the larger seasonal profile samples. In profile, overall decreases in 18O and 13C were observed, supporting evidence that positive temperature anomalies caused the bleaching event and reflecting the loss of zooxanthellae photosynthesis.  相似文献   

8.
The stable isotopic composition of soil water is controlled by precipitation inputs, antecedent conditions, and evaporative losses. Because transpiration does not fractionate soil water isotopes, the relative proportions of evaporation and transpiration can be estimated using a simple isotopic mass balance approach. At our site in the shortgrass steppe in semi-arid northeastern Colorado, 18O values of soil water were almost always more enriched than those of precipitation inputs, owing to evaporative losses. The proportion of water lost by evaporation (E/ET) during the growing season ranged from nil to about 40% (to >90% in the dormant season), and was related to the timing of precipitation inputs. The sum of transpiration plus evaporation losses estimated by isotopic mass balance were similar to actual evapotranspiration measured from a nearby Bowen ratio system. We also investigated the evapotranspiration response of this mixed C3/C4 grassland to doubled atmospheric [CO2] using Open-Top Chambers (OTC). Elevated atmospheric [CO2] led to increased soil-water conservation via reduced stomatal conductance, despite greater biomass growth. We used a non-invasive method to measure the 18O of soil CO2 as a proxy for soil water, after establishing a strong relationship between 18O of soil CO2 from non-chambered control (NC) plots and 18O of soil–water from an adjacent area of native grassland. Soil–CO2 18O values showed significant treatment effects, particularly during a dry summer: values in ambient chambers (AC) were more enriched than in NC and elevated chamber (EC) plots. During the dry growing season of 2000, transpiration from the EC treatment was higher than from AC and lower than from NC treatments, but during 2001, transpiration was similar on all three treatments. Slightly higher evaporation rates from AC than either EC or NC treatments in 2000 may have resulted from increased convection across the soil surface from the OTC blowers, combined with lower biomass and litter cover on the AC treatment. Transpiration-use efficiency, or the amount of above-ground biomass produced per mm water transpired, was always greatest on EC and lowest on NC treatments.  相似文献   

9.
Some of the oxygen produced during oxygenic photosynthesis is consumed but little is known about the extent of the processes involved. We measured the 17O/16O and 18O/16O ratios in O2 produced by certain marine and freshwater phytoplankton representing important groups of primary producers. When the cells were performing photosynthesis under very low dissolved oxygen concentrations (<3 μM), we observed significant enrichment in both 18O and 17O with respect to the substrate water. The difference in δ18O between O2 and water was about 4.5, 3, 5.5, and 7‰ in the diatom Phaeodactylum tricornutum, Nannochloropsis sp. (Eustigmatophyceae), the coccolithophore Emiliania huxleyi and the green alga Chlamydomonas reinhardtii, respectively. The difference in δ17O was about 0.52 that of δ18O. As explained, the observed enrichments most probably stem from considerable oxygen consumption during photosynthesis even when major O2-consuming reactions such as photorespiration were minimized. These enrichments increased linearly with rising O2 levels but with different δ17O/δ18O slopes for the various organisms, suggesting engagements of different O2-consuming reactions with rising O2 levels. Consumption of O2 may be important for energy dissipation during photosynthesis. The isotope enrichment observed here, not accounted for in earlier assessments, closes an important gap in our understanding of the difference between the isotopic compositions of atmospheric oxygen and that of seawater, i.e., the Dole effect.  相似文献   

10.
Helliker BR 《Plant physiology》2011,155(4):2096-2107
Previous theoretical work showed that leaf-water isotope ratio (δ18OL) of Crassulacean acid metabolism epiphytes was controlled by the δ18O of atmospheric water vapor (δ18Oa), and observed δ18OL could be explained by both a non-steady-state model and a “maximum enrichment” steady-state model (δ18OL-M), the latter requiring only δ18Oa and relative humidity (h) as inputs. δ18OL, therefore, should contain an extractable record of δ18Oa. Previous empirical work supported this hypothesis but raised many questions. How does changing δ18Oa and h affect δ18OL? Do hygroscopic trichomes affect observed δ18OL? Are observations of changes in water content required for the prediction of δ18OL? Does the leaf need to be at full isotopic steady state for observed δ18OL to equal δ18OL-M? These questions were examined with a climate-controlled experimental system capable of holding δ18Oa constant for several weeks. Water adsorbed to trichomes required a correction ranging from 0.5‰ to 1‰. δ18OL could be predicted using constant values of water content and even total conductance. Tissue rehydration caused a transitory change in δ18OL, but the consequent increase in total conductance led to a tighter coupling with δ18Oa. The non-steady-state leaf water models explained observed δ18OL (y = 0.93*x − 0.07; r2 = 0.98) over a wide range of δ18Oa and h. Predictions of δ18OL-M agreed with observations of δ18OL (y = 0.87*x − 0.99; r2 = 0.92), and when h > 0.9, the leaf did not need to be at isotopic steady state for the δ18OL-M model to predict δ18OL in the Crassulacean acid metabolism epiphyte Tillandsia usneoides.Tropical and subtropical epiphytic higher plants have long been a curiosity to plant physiologists because of the multiple and unique constraints on physiology that the epiphytic lifeform represents (Mez, 1904; Benzing, 1970; Medina and Troughton, 1974). While the competition for light has no doubt led the plants to the tree tops, the concomitant loss of roots effectively removed the plants from an environment of abundant rainfall to one of arid conditions (Griffiths et al., 1986; Smith et al., 1986a; Winter and Smith, 1996). In a sense, the plants shifted biomes by immigrating vertically instead of horizontally. This shift led to morphological adaptations like the development of modified leaf hairs to absorb water and the formation of tanks via the overlapping of leaf bases as well as physiological adaptations such as maintaining high leaf osmotic potential and, perhaps most notably, the evolution of the CO2-concentrating mechanism known as Crassulacean acid metabolism (CAM) photosynthesis (Medina and Troughton, 1974; Benzing et al., 1976; Griffiths and Smith, 1983; Smith et al., 1986b; Martin and Schmitt, 1989; Martin et al., 2004; Ohrui et al., 2007). This unique physiology of tropical and subtropical CAM epiphytes also presents an intriguing contrast to the general way we view oxygen isotope ratios (δ18O) in plant water and organic material (Helliker and Griffiths, 2007; Helliker and Noone, 2009).The enrichment of 18O in leaf water during plant transpiration leads to a suite of physiology-based tracers that inform us of current and past plant-environment interactions. Leaf water δ18O (δ18OL) labels CO2 and allows for partitioning of the gross components of net CO2 flux from ecosystem to regional scales (Yakir and Wang, 1996; Ciais et al., 1997; Styles et al., 2002; Cuntz et al., 2003; Ogee et al., 2004; Helliker et al., 2005). The production of isotopically distinct O2 by photosynthesis allows for global-scale estimates of productivity over millennia (Guy et al., 1993; Luz et al., 1999; Hoffmann et al., 2004). The isotopic record of δ18OL in leaf and tree ring cellulose allows for the reconstruction of growth environment and/or physiological responses to that growth environment (Epstein et al., 1977; Anderson et al., 1998; Switsur and Waterhouse, 1998; Barbour and Farquhar, 2000; Barbour et al., 2000; Roden and Ehleringer, 2000; Ferrio and Voltas, 2005; Poussart and Schrag, 2005; Helliker and Richter, 2008). Much of the error associated with the above δ18O applications could be decreased with better estimates of the isotope ratio of atmospheric water vapor (δ18Oa), a primary control on leaf-water 18O enrichment (Farquhar and Cernusak, 2005; Helliker and Griffiths, 2007), and the study of δ18OL in CAM epiphytes may offer better estimates of δ18Oa through time and space (Helliker and Griffiths, 2007).The epiphyte, and specifically the CAM vascular epiphyte, offers an extreme in the way δ18O in plant water and organic material is viewed. For most vascular plants, δ18OL is controlled by a balance of soil water δ18O and δ18Oa; soil water comes into the leaves via the root system, and atmospheric vapor diffuses into the leaf through open stomata. The balance of the two water sources is determined by relative humidity (h): if h is high, then δ18OL is controlled proportionally more by δ18Oa; if h is unity, then δ18OL is at equilibrium with δ18Oa. If these plants happen to lose water during a humid night, then leaf water exchange can lead to control by, and possibly equilibrium with, δ18Oa (Lai et al., 2008). Typically, the dramatic increase of transpiration during the daytime moves the δ18OL away from equilibrium with δ18Oa. Hence, the nocturnal movement of δ18OL toward equilibrium with δ18Oa in most vascular plants is not likely recorded in plant organic material and is only important to the isotopic composition of CO2 during nocturnal respiration (Cernusak et al., 2004). In the CAM epiphyte, rainwater and dewfall rehydrate epiphyte tissue, and due to stomata being open at night, when h is near unity, rehydration water is quickly exchanged with atmospheric water vapor. This leads to the situation, hypothesized by Helliker and Griffiths (2007), where rainfall and dewfall (which are in equilibrium with δ18Oa; Gat, 1996) control tissue water status, and the δ18O of this tissue water is continuously controlled by δ18Oa through subsequent nocturnal vapor exchange. Because stomata are typically closed during the day, organic material synthesized by the photosynthetic carbon reduction cycle should obtain the δ18O signature of δ18Oa-controlled leaf water (Helliker and Griffiths, 2007). It is this situation that yields new applications for δ18O that are unique to epiphytes: the reconstruction of physiological responses to the epiphytic growth habit (Reyes-Garcia et al., 2008) and the reconstruction of δ18Oa (Helliker and Griffiths, 2007). It was recently shown through empirical and theoretical work that lichen thalli obtain equilibrium with δ18Oa even under nonsaturating conditions (Hartard et al., 2009). The lichen system of equilibrium with δ18Oa represents an important contrast to that of CAM epiphytes, and this is discussed later. The theoretical work presented by Hartard et al. (2009), however, is extremely helpful for the interpretation of isotopic signals in CAM epiphytes.Helliker and Griffiths (2007) developed the theoretical underpinnings for δ18Oa as a control on CAM epiphyte δ18OL. They extended the thoughts of Farquhar and Cernusak (2005) to show that at the high nocturnal h experienced by CAM epiphytes, gross exchange fluxes of water vapor into the leaf from the atmosphere can be several times the transpirational flux out of the leaf. Model simulations showed that constant environmental conditions and continuous water loss led to an isotopic steady state where both δ18OL and transpired water (δ18OE) converged to a single isotopic value controlled by the exchange of δ18Oa and described by the following “maximum enrichment” equation (Farquhar and Gan, 2003; Helliker and Griffiths, 2007; Hartard et al., 2009):where the steady-state δ18OL (RL) is determined solely by δ18Oa (Ra), h, the temperature-dependent equilibrium fractionation factor α*, and the balance of the ratio of diffusivities of light to heavy water molecules through the stomata and through the leaf boundary layer, αK (Flanagan et al., 1991; Farquhar and Lloyd, 1993; Farquhar and Cernusak, 2005). The non-steady-state analytical solution for δ18OL developed by Hartard et al. (2009; Eq. 6 in “Materials and Methods,” hereafter referred to as the Hartard-Cuntz solution) clearly demonstrates that, through time in the nonsteady state, δ18OL is continuously moving toward the value of Rl-M18Ol-M). In a similar manner, the simulations of Helliker and Griffiths (2007) suggested that even at values of nocturnal h of 0.8, which is low for a CAM epiphyte (Garth, 1964; Smith et al., 1986b), δ18OL was controlled almost entirely by the isotope ratio of atmospheric water vapor and the maximum enrichment equation above would ultimately predict δ18OL. The simulations also showed that the approach to steady state was faster and occurred through less water loss as h increased. In general, their simulations showed that δ18OL approached the steady-state value, or δ18Ol-M, much faster than δ18OE, and this yields an important prediction that is tested by our study: the leaf does not need to be at full isotopic steady state for δ18OL to be at or near the steady-state, maximum enrichment value (δ18Ol-M).There are many unresolved questions as to the agreement between models and observations that underpin the efficacy of using δ18OL to reconstruct δ18Oa. The empirical work of Helliker and Griffiths (2007) did show support for the modeling exercises, but only over a narrow range of conditions. Their experimental setup was limited, as it was developed to demonstrate a preliminary proof of concept. Also, the invariably high h and noisy δ18Oa (±1.2‰) of Helliker and Griffiths (2007) could lead one to erroneously conclude that δ18OL was in direct equilibrium with δ18Oa even when h was less than unity, which should not be the case. The primary controls of CAM epiphyte δ18OL are h and δ18Oa, and the manner in which changes in δ18Oa and h affect δ18OL through time must be assessed. Like many CAM epiphytes, the study species Tillandsia usneoides has a heavy covering of hygroscopic trichomes. While the water adsorbed to these trichomes does not lead to water uptake by living cells at subsaturating conditions (Martin and Schmitt, 1989), the water is inevitably sampled and extracted for δ18OL determination. Hence, the isotopic offset caused by these trichomes must be determined. Total plant conductance to water loss (gtot; stomatal and boundary layer conductance) controls the rate of exchange of tissue water with water vapor and the relative water content (RWC) of the plant. Therefore, the effect of changes in RWC on δ18OL, through both water loss and rehydration, must be determined. Previous work has shown that changes in water volume are not important to predictions of δ18OL (Cernusak et al., 2002; Cuntz et al., 2007), and a similar finding in CAM epiphytes could greatly simplify our interpretation of observed δ18OL. In summary, the goal of this study was to construct controlled-environment experiments to assess how well observed δ18OL could be predicted over a range of conditions through both steady-state and non-steady-state approaches of Helliker and Griffiths (2007) and Hartard et al. (2009).  相似文献   

11.
The effect of irradiance on the rate of net photosynthesis was measured for mature leaves of coffee grown under five levels of radiation from 100% to 5% daylight. The rate of light-saturated photosynthesis per unit leaf area (PNmax) increased from 2 mol CO2 m-2 s-1 under 5% daylight to 4.4 mol CO2 m-2 s-1 under 100% daylight. The photon flux density (PAR, photosynthetically active radiation) needed for 50% saturation of photosynthesis, as well as the light compensation point, also increased with increasing levels of irradiation during growth. The quantum efficiency of photosynthesis (), measured by the initial slope of the photosynthetic response to increasing irradiance, was greater under shaded growth conditions. The rate of dark respiration was greatest for plants grown in full daylight. On the basis of the increase in the quantal efficiency of photosynthesis and the low light compensation point when grown under shaded conditions, coffee shows high shade adaptation. Plants adjusted to shade by an increased ability to utilize short-term increases in irradiance above the level of the growth irradiance (measured by the difference between photosynthesis at the growth irradiance, PNg, and PNmax).  相似文献   

12.
H. Fock  K. Klug  D. T. Canvin 《Planta》1979,145(3):219-223
Using an open gas-exchange system, apparent photosynthesis, true photosynthesis (TPS), photorespiration (PR) and dark respiration of sunflower (Helianthus annuus L.) leaves were determined at three temperatures and between 50 and 400 l/l external CO2. The ratio of PR/TPS and the solubility ratio of O2/CO2 in the intercellular spaces both decreased with increasing CO2. The rate of PR was not affected by the CO2 concentration in the leaves and was independent of the solubility ratio of oxygen and CO2 in the leaf cell. At photosynthesis-limiting concentrations of CO2, the ratio of PR/TPS significantly increased from 18 to 30°C and the rate of PR increased from 4.3 mg CO2 dm-2 h-1 at 18°C to 8.6 mg CO2 dm-2 h-1 at 30°C. The specific activity of photorespired CO2 was CO2-dependent but temperature-independent, and the carbon traversing the glycolate pathway appeared to be derived both from recently fixed assimilate and from older reserve materials. It is concluded that PR as a percentage of TPS is affected by the concentrations of O2 and CO2 around the photosynthesizing cells, but the rate of PR may also be controlled by other factors.Abbreviations APS apparent photosynthesis (net CO2 uptake) - PR photorespiration (CO2 evolution in light) - RuBP ribulose-1,5-bisphosphate - TPS true photosynthesis (true CO2 uptake)  相似文献   

13.
Recovery from photoinhibition of photosynthesis in intact Lemna gibba was studied in presence of the protein synthesis inhibitors chloramphenicol and cycloheximide. Exposure to an irradiance of 1000 mol m-2s-1 in N2 for 90 min induced 80% photoinhibition. The plants recovered photosynthesis when transfered to normal irradiances (210 mol m-2s-1) and air. Chloramphenicol added to the medium was taken up by the plant and reduced photosynthesis slightly. Recovery from photoinhibition was more inhibited than photosynthesis. Cycloheximide was also taken up by the plants and reduced synthesis of light harvesting chlorophyll protein: however, neither photosynthesis nor recovery were much affected. Synthesis of 32-kD chloroplast protein during recovery was inhibited by chloramphenicol, but not by cycloheximide. Synthesis of 32-kD protein was enhanced by 20–210 mol m-2s-1 light. The results support the hypothesis that synthesis of 32-kD protein is important for recovery of photosynthesis after photoinhibition.  相似文献   

14.
Klaus Winter 《Oecologia》1981,48(1):100-106
Summary Measurements of 13C values were used as a diagnostic test for the possible occurrence of C4 photosynthesis in 175 species of the Chenopodiaceae and 18 species of the genus Calligonum (Polygonaceae) from deserts of the Middle East and USSR. Eighty percent of the Chenopodiaceae (predominantly members of the genera Aellenia, Anabasis, Haloxylon, Salsola and Suaeda) and all species of the genus Calligonum showed C4-like 13C values. Several features of these plants disclose some new facets of C4 photosynthesis. Some of the Haloxylon and Calligonum species are trees or tall shrubs and in Middle Asia are dominant members of plant communities characterized by high biomass. Many C4 species of the Chenopodiaceae and Polygonaceae in their natural, Middle and Central Asian desert habitats, experience temperatures far below the freezing point for a long period of the year. Several of these C4 species are of considerable economic value.  相似文献   

15.
A phylloquinone molecule (2-methyl, 3-phytyl, 1, 4-naphthoquinone) occupies the A1 binding site in photosystem 1 particles from Synechocystis sp. 6803. In menB mutant photosystem 1 particles from the same species, plastoquinone-9 occupies the A1 binding site. By incubation of menB mutant photosystem 1 particles in the presence of phylloquinone, it was shown in another study that phylloquinone will displace plastoquinone-9 in the A1 binding site. We describe the reconstitution of unlabeled (16O) and 18O-labeled phylloquinone back into the A1 binding site in menB photosystem 1 particles. We then produce time-resolved Fourier transform infrared (FTIR) difference spectra for these menB photosystem 1 particles that contain unlabeled and 18O-labeled phylloquinone. By specifically labeling only the phylloquinone oxygen atoms we are able to identify bands in FTIR difference spectra that are due to the carbonyl (CO) modes of neutral and reduced phylloquinone. A positive band at 1494 cm−1 in the FTIR difference spectrum is found to downshift 14 cm−1 and decreases in intensity on 18O labeling. Vibrational mode frequency calculations predict that an antisymmetric vibration of both CO groups of the phylloquinone anion should display exactly this behavior. In addition, phylloquinone that has asymmetrically hydrogen bonded carbonyl groups is also predicted to display this behavior. The positive band at 1494 cm−1 in the FTIR difference spectrum is therefore due to the antisymmetric vibration of both CO groups of one electron reduced phylloquinone. Part of a negative band at 1654 cm−1 in the FTIR difference spectrum downshifts 28 cm−1 on 18O labeling. Again, vibrational mode frequency calculations predict this behavior for a CO mode of neutral phylloquinone. The negative band at 1654 cm−1 in the FTIR difference spectrum is therefore due to a CO mode of neutral phylloquinone. More specifically, calculations on a phylloquinone model molecule with the C4O group hydrogen bonded predict that the 1654 cm−1 band is due to the non hydrogen bonded C1O mode of neutral phylloquinone.  相似文献   

16.
Summary Characteristics of inorganic carbon assimilation by photosynthesis in seawater were investigated in six species of the Fucales (five Fucaceae, one Cystoseiraceae) and four species of the Laminariales (three Laminariaceae, one Alariaceae) from Arbroath, Scotland. All of the algae tested could photosynthesise faster at high external pH values than the uncatalysed conversion of HCO 3 - to CO2 can occur, i.e. can use external HCO 3 - . They all had detectable extracellular carbonic anhydrase activity, suggesting that HCO 3 - use could involve catalysis of external CO2 production, a view supported to some extent by experiments with an inhibitor of carbonic anhydrase. All of the algae tested had CO2 compensation concentrations at pH 8 which were lower than would be expected from diffusive entry of CO2 supplying RUBISCO as the initial carboxylase, consistent with the operation of energized entry of HCO 3 - and / or CO2 acting as a CO2 concentrating mechanism. Quantitative differences among the algae examined were noted with respect to characteristics of inorganic C assimilation. The most obvious distinction was between the eulittoral Fucaceae, which are emersed for part of, or most of, the tidal cycle, and the other three families (Cystoseiraceae, Laminariaceae, Alariaceae) whose representatives are essentially continually submersed. The Fucaceae examined are able to photosynthesise at high pH values, and have lower CO2 compensation concentrations, and lower K1/2 values for inorganic C use in photosynthesis, at pH 8, than the other algae tested. Furthermore, the Fucaceae are essentially saturated with inorganic C for photosynthesis at the normal seawater concentration at pH 8 and 10°C. These characteristics are consistent with the dominant role of a CO2 concentrating mechanism in CO2 acquisition by these plants. Other species tested have characteristcs which suggest a less effective HCO 3 - use and CO2 concentrating mechanism, with the Laminariaceae being the least effective; unlike the Fucaceae, photosynthesis by these algae is not saturated with inorganic C in normal seawater. Taxonomic and ecological implications of these results are considered in relation to related data in the literature.  相似文献   

17.
Attached leaves of sunflower (Helianthus annuus L.) were exposed to 14CO2 during steady-state photosynthesis for 2 to 30 min in 345 l/l CO2 and 21% O2 at 29° C and a light intensity of 1300 E m-2s-1. Glycolic acid was extracted with water and diethyl ether, and was determined in the aqueous residue by high-pressure liquid column chromatography. The relative specific radioactivity of the glycolic acid synthesized during photosynthesis reached about 100% after 30 min of photosynthesis and was almost equal to that of the CO2 evolved during photorespiration, their ratio at all times being nearly one. These results provide strong in-vivo evidence that the glycolic acid is the substrate for CO2 evolved by sunflower leaves in light.  相似文献   

18.
19.
20.
Effects of environmental conditions on isoprene emission from live oak   总被引:12,自引:0,他引:12  
Live-oak plants (Quercus virginiana Mill.) were subjected to various levels of CO2, water stress or photosynthetic photon flux density to test the hypothesis that isoprene biosynthesis occurred only under conditions of restricted CO2 availability. Isoprene emission increases as the ambient CO2 concentration decreased, independent of the amount of time that plants had photosynthesized at ambient CO2 levels. When plants were water-stressed over a 4-d period photosynthesis and leaf conductance decreased 98 and 94%, respectively, while isoprene emissions remained constant. Significant isoprene emissions occurred when plants were saturated with CO2, i.e., below the light compensation level for net photosynthesis (100 mol m-2 s-1). Isoprene emission rates increased with photosynthetic photon flux density and at 25 and 50 mol m-2 s-1 were 7 and 18 times greater than emissions in the dark. These data indicate that isoprene is a normal plant metabolite and not — as has been suggested — formed exclusively in response to restricted CO2 or various stresses.Abbreviation PPFD photosynthetic photon flux density  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