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.
Künnis  Kai 《Hydrobiologia》1997,363(1-3):253-260
Microbial degradation of algal detritus was studied experimentally usingthe diatom Skeletonema costatum prekilled culture as a substrate for themarine microbial community. The qualitative and quantitative changes in themicrobial community and the algal detritus structure were followed during 11days of incubation at the water temperature of 20°C. Most of thebacterial parameters (epifluorescence microscopy counts of free-living andattached cells, mean cell volume, biomass and productivity) were the highestafter 24 h from the algal detritus addition to the microbial community. Thebacterial peak (3.6×106 cells ml™1,biomass 3.1 μgC ml™1, net production 66×4 ngC ml™1h™1) was followed by a precipitous increaseof homogenous nanoflagellate population with a maximum number of2.6×105 cells ml™1, which in turn declinedquickly after ciliates appeared in the community. The bacterial production,initiated by the supplement of algal detritus, was totally ingested bymicrozoans within 3 days. Changes of the structure of marine bacterialassemblage and relative increase of the amount of attached bacteria duringthe period of massive development of nanoflagellates emphasized theimportance of small flagellates predation on the free-living average size(cell volume 0.1–0.2 μm3) pelagic bacteria. After11 days of incubation on algal detritus the initial bacterial assemblage wasreplaced by the mixed succession of bacteria, flagellates and ciliates at aproportion of 1000 : 1.5 : 0.2. The stabilization of microbialcommunity and changes in algal detritus structure allow to expect thatpelagic microbial utilization of Skeletonema costatum bloom(53×106 cells l™1), could be finishedwithin a period of 8–11 days at the summer water temperature. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

2.
Four surveys were carried out in August 2001, and in August, September and October 2002 in the southern Yellow Sea, China. Bacterioplankton abundance and biomass were quantified along the transects from Qingdao, China to the Korean island of Jeju. In this paper, the author describes the distribution of bacterioplankton and predatory pressures, and also discusses the impact of Yellow Sea Cold Water Mass (YSCWM) on the distribution of bacterioplankton. The range of Synechococcus biomass was 0.78–33.49 mg C/m3 (on average 6.26 mg C/m3), and heterotrophic bacteria biomass was 1.58–21.25 mg C/m3 (on average 5.79 mg C/m3, n=197). In the vertical direction, the distribution order of Synechococcus biomass was mid-layer>surface>bottom, and the one of heterotrophic bacteria biomass was surface>mid-layer>bottom. The contribution of Synechococcus to the total phytoplankton biomass (SB/PB) ranged from 2% to 99% (on average 42.5%), and the ratio of heterotrophic bacterial biomass to phytoplankton biomass (HBB/PB) was from 5% to 637% (on average 85%). Daily variations of bacterioplankton biomass at anchor stations showed that the maximal Synechococcus biomass was 8.8 times the minimum, and it was 2.8 times in heterotrophic bacterial biomass. The distribution of bacterioplankton accorded with temperature and salinity, and the minimum of bacterioplankton biomass occurred in the YSCWM waters. At the same time, the microzooplankton (20–200 μm) preyed on Synechococcus in these waters, with the ingestion rate of about 0.20–0.42/d.  相似文献   

3.
Proposals for macrophyte restoration in eutrophic coastal lagoons   总被引:2,自引:2,他引:0  
Based on the comparison of environmental requirements forRuppia cirrhosa andPotamogeton pectinatus growth, macrophyteversus phytoplankton biomass and production features, and differences in hydrological and nutrient balances between Tancada lagoon (where macrophytes form dense beds) and Enca?izada lagoon (with no macrophytes at all), several proposals for macrophyte restoration are presented. The highest photosynthetic efficiency ofR. cirrhosa takes place at high irradiance and it grows over a wide range of salinity.P. pectinatus is better adapted to lower light intensity and salinity thanRuppia. R. cirrhosa transplanted from Tancada to Enca?izada was successful in enclosures, where light availability increased (μ=0.013 cm−1), but not in open waters where light extinction coefficient was 0.032 cm−1. Phytoplankton biomass (0.11–2.15 g C m−2) is much lower than macrophyte biomass (16–200 g C m−2) in Tancada lagoon. However, phytoplankton production (165 g C m−2 yr−1 in Tancada, 480 g C m−2 yr−1 in Enca?izada) is the same order of magnitude as macrophyte production (244–467 g C m−2 yr−1). Turnover rates are 0.3–0.9 day−1 for phytoplankton and 1.2–2.5 yr−1 for macrophytes. Phytoplankton and inorganic particles are responsible for high turbidity of the water in Enca?izada lagoon. Phytoplankton blooms in Enca?izada lagoon are supported by high freshwater inflows from rice field drains from May to November. The Qs (seawater discharge)/QF (Freshwater discharge) ratios are, respectively, 0.24 and 0.48, which denotes a higher seawater influence in Tancada than in Enca?izada lagoon. Decreasing freshwater inputs to Enca?izada lagoon both in May and November thus allowing greater inputs of sea water, is proposed as the most effective way to restore this eutrophic coastal lagoon. The objective being to reduce nutrient loadings to the lagoon and phytoplankton in order to favour macrophyte re-colonization.  相似文献   

4.
Two bacterial species isolated using enrichment culture techniques from the topsoil of a main refuse dumpsite in Nigeria were assessed for their dehalogenation potentials. The bacterial isolates were identified as belonging to the Bacillus and Pseudomonas genera. Axenic cultures of the isolates utilized monochloroacetic acid (MCA), trichloroacetic acid (TCA), trichloromethane (CHCl3) and tetrachloromethane (CCl4) as the sole source of carbon for growth up to a final substrate concentration of 0.1% (w/v). The mean generation times of the isolates in all the growth media ranged significantly (P<0.05) from 2.41 to 10.04 h and were generally higher than that observed in glucose medium (1.46–1.51 h). The numbers of the chloride atoms in the different organochlorides were negatively correlated with the ability of the organisms to degrade the compounds. Dehalogenase specific activities of the cell-mediated cultures ranged from 0.1 to 0.96 μg ml–1 chloride release (mg protein)–1 h–1 and were significantly (P <0.05) higher than that of the cell-free extract [0.09–0.8 μg ml–1 chloride release (mg protein)–1 h–1]. The optimal pH of the dehalogenase activity was found to be 8.0, and the optimal temperature was between 30 and 35 °C. Electronic Publication  相似文献   

5.
During two intensive field campaigns in summer and autumn 2004 nitrogen (N2O, NO/NO2) and carbon (CO2, CH4) trace gas exchange between soil and the atmosphere was measured in a sessile oak (Quercus petraea (Matt.) Liebl.) forest in Hungary. The climate can be described as continental temperate. Fluxes were measured with a fully automatic measuring system allowing for high temporal resolution. Mean N2O emission rates were 1.5 μg N m−2 h−1 in summer and 3.4 μg N m−2 h−1 in autumn, respectively. Also mean NO emission rates were higher in autumn (8.4 μg N m−2 h−1) as compared to summer (6.0 μg N m−2 h−1). However, as NO2 deposition rates continuously exceeded NO emission rates (−9.7 μg N m−2 h−1 in summer and −18.3 μg N m−2 h−1 in autumn), the forest soil always acted as a net NO x sink. The mean value of CO2 fluxes showed only little seasonal differences between summer (81.1 mg C m−2 h−1) and autumn (74.2 mg C m−2 h−1) measurements, likewise CH4uptake (summer: −52.6 μg C m−2 h−1; autumn: −56.5 μg C m−2 h−1). In addition, the microbial soil processes net/gross N mineralization, net/gross nitrification and heterotrophic soil respiration as well as inorganic soil nitrogen concentrations and N2O/CH4 soil air concentrations in different soil depths were determined. The respiratory quotient (ΔCO2 resp ΔO2 resp−1) for the uppermost mineral soil, which is needed for the calculation of gross nitrification via the Barometric Process Separation (BaPS) technique, was 0.8978 ± 0.008. The mean value of gross nitrification rates showed only little seasonal differences between summer (0.99 μg N kg−1 SDW d−1) and autumn measurements (0.89 μg N kg−1 SDW d−1). Gross rates of N mineralization were highest in the organic layer (20.1–137.9 μg N kg−1 SDW d−1) and significantly lower in the uppermost mineral layer (1.3–2.9 μg N kg−1 SDW d−1). Only for the organic layer seasonality in gross N mineralization rates could be demonstrated, with highest mean values in autumn, most likely caused by fresh litter decomposition. Gross mineralization rates of the organic layer were positively correlated with N2O emissions and negatively correlated with CH4 uptake, whereas soil CO2 emissions were positively correlated with heterotrophic respiration in the uppermost mineral soil layer. The most important abiotic factor influencing C and N trace gas fluxes was soil moisture, while the influence of soil temperature on trace gas exchange rates was high only in autumn.  相似文献   

6.
Changes in leaf physiology with tree age and size could alter forest growth, water yield, and carbon fluxes. We measured tree water flux (Q) for 14 ponderosa pine trees in two size classes (12 m tall and ∼40 years old, and 36 m tall and ∼ 290 years old) to determine if transpiration (E) and whole-tree conductance (g t) differed between the two sizes of trees. For both size classes, E was approximately equal to Q measured 2 m above the ground: Q was most highly correlated with current, not lagged, water vapor pressure deficit, and night Q was <12% of total daily flux. E for days 165–195 and 240–260 averaged 0.97 mmol m–2 (leaf area, projected) s–1 for the 12-m trees and 0.57 mmol m–2 (leaf area) s–1 for the 36-m trees. When photosynthetically active radiation (I P) exceeded the light saturation for photosynthesis in ponderosa pine (900 μmol m–2 (ground) s–1), differences in E were more pronounced: 2.4 mmol m–2 (leaf area) s–1 for the 12-m trees and 1.2 mmol m–2 s–1 for the 36-m trees, yielding g t of 140 mmol m–2 (leaf area) s–1 for the 12-m trees and 72 mmol m–2 s–1 for the 36-m trees. Extrapolated to forests with leaf area index =1, the 36-m trees would transpire 117 mm between 1 June and 31 August compared to 170 mm for the 12-m trees, a difference of 15% of average annual precipitation. Lower g t in the taller trees also likely lowers photosynthesis during the growing season. Received: 19 April 1999 / Accepted: 23 March 2000  相似文献   

7.
Water status and gas exchange of beech (Fagus sylvatica L.) and sessile oak [Quercus petraea (Mattuschka) Liebl.] were studied in a mixed stand in the Montejo de la Sierra forest (central Iberian Peninsula), one of the southernmost locations of both species in Europe. Gas exchange and water potential were measured in leaves at different canopy levels over several days in two growing seasons. The daily variation pattern was established with the measurements of three selected dates per year, representative of the soil moisture content situations in early, mid- and late summer. A similar daily time course of leaf water potential was found for the two species. Nevertheless, beech showed a most noticeable decrease of water potential at midmorning and maintained lower leaf water potential than oak in the early afternoon. In 1994 the sessile oak saplings showed higher values of predawn water potential (Ψpd) than beech at the end of summer, when soil moisture content was lowest (20 cm depth). Beech showed a significantly lower net assimilation rate (A) than sessile oak for leaves under the same PPFD. Maximum net photosynthesis values (A max) for beech and sessile oak on sunny leaves were 10.1±0.4 μmol m–2 s–1 and 17.8±1.7 μmol m–2 s–1 respectively, and those for water vapour stomatal conductance (g wv) were 265±31 mmol m–2 s–1 and 438±74 mmol m–2 s–1. Differences in A and g wv between the two species were maintained throughout the day on all measurement dates. No clear relationship was found between water status of saplings and stomata performance; there was only a negative correlation between Ψpd and g wvmid in beech. Nevertheless, a significant response to the air vapour pressure gradient between leaf and air was translated into stomata closure on an hourly basis, more intensively in beech. Received: 4 March 1999 / Accepted: 21 December 1999  相似文献   

8.
The sulfur cycle of Ebro Delta microbial mats was studied in order to determine sulfide production and sulfide consumption. Vertical distribution of two major functional groups involved in the sulfur cycle, anoxygenic phototrophic bacteria and dissimilatory sulfate-reducing bacteria (SRB), was also studied. The former reached up to 2.2×108 cfu cm–3 sediment in the purple layer, and the latter reached about 1.8×105 SRB cm–3 sediment in the black layer. From the changes in sulfide concentrations under light-dark cycles it can be inferred that the rate of H2S production was 6.2 μmol H2S cm–3 day–1 at 2.6 mm, and 7.6 μmol H2S cm–3 day–1 at 6 mm. Furthermore, sulfide consumption was also assessed, determining rates of 0.04, 0.13 and 0.005 mmol l–1 of sulfide oxidized at depths of 2.6, 3 and 6 mm, respectively. Electronic Publication  相似文献   

9.
Geographic patterns of intraspecific variations in traits related to photosynthesis and biomass were examined in two separate common garden experiments using seed collected from 26 Sitka alder (Alnus sinuata Rydb.) and 18 paper birch (Betula papyrifera Marsh.) populations from climatically diverse locations in British Columbia, Canada. Exchange rates of carbon dioxide and water vapour were measured on 2-year-old seedlings to determine the maximum net instantaneous photosynthetic rate, mesophyll conductance, stomatal conductance, and photosynthetic water use efficiency. Height, stem diameter, root and shoot dry mass and fall frost hardiness data were also obtained. Mean population maximum photosynthetic rate ranged from 10.35 to 14.57 μmol CO2 m–2 s–1 in Sitka alder and from 14.76 to 17.55 μmol CO2 m–2 s–1 in paper birch. Based on canonical correlation analyses, populations from locations with colder winters and shorter (but not necessarily cooler) summers had higher maximum photosynthetic rates implying the existence of an inverse relationship between leaf longevity and photosynthetic capacity. Significant canonical variates based on climatic variables derived for the seed collection sites explained 58% and 41% of variation in the rate of photosynthesis in Sitka alder and paper birch, respectively. Since growing season length is reflected in date of frost hardiness development, an intrinsic relationship was found between photosynthetic capacity and the level of fall frost hardiness. The correlation was particularly strong for paper birch (r=–0.77) and less strong for Sitka alder (r=–0.60). Mean population biomass accumulation decreased with increased climate coldness. These patterns may be consequential for evaluation of the impact of climate change and extension of the growing season on plant communities. Received: 12 July 1999 / Accepted: 24 November 1999  相似文献   

10.
Summary Chickpea was micropropagated by axillary shoot proliferation (ASP) and modified single node culture (MSNC) methods. Maximum propagule proliferation occurred on Murashige and Skoog (MS) medium enriched with 1–10 μM N6-benzyladenine and 0.01 μM α-naphthaleneacetic acid. The propagules were rooted on MS medium containing 1 μM 3-indolebutyric acid and B5 vitamins. Regenerated plants were fertile and phenotypically similar to control plants grown from seed. The MSNC method was four times more efficient than the ASP method in terms of the number of plants produced per explant.  相似文献   

11.
Xie  Ping  Iwakuma  Toshio  Fujii  Koichi 《Hydrobiologia》1998,368(1-3):83-90
Population dynamics of Chaoborus flavicans larvae of various instars was studied from November 1986 to December 1987 in a eutrophic, fish-free pond, Japan. First and 2nd instar larvae were observed from late April to late October, indicating a reproductive period of about half a year. C. flavicans overwintered in the 4th instar larvae. In water column samples, total density of all instars was 680–23680 m-2, and pupal density 0–2600 m-2; larvae of the 1st, 2nd, and 3rd instars showed 5–6 density peaks in 1987, suggesting that 5–6 generations occur during a year (peaks of the 4th instar larvae were not clear, probably due to their longer development than those of younger instars). In sediment samples, no 1st and 2nd instar larvae were found, 3rd instar larvae were found occasionally but density of the 4th instar larvae was 280–18600 m-2, and pupal density varied between 0–502 m-2. Fouth instar larvae accumulated in sediment in the cold season and in the water column in the warm season; high temperature and low oxygen concentration were the most important factors limiting the distribution of larvae in the sediment in summer in the NIES pond. The dry weight of total C. flavicans larvae was 0.08–4.2 g m-2 in sediment samples and 24–599 μg l-1 (0.10–2.40 g m-2) in water column samples. Comparisons of maximum densities in the NIES pond in different years and in waters of different trophic status show that density is generally higher in eutrophic than in oligotrophic habitats. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

12.
This study was undertaken in summer on fully expanded leaves of Avicennia marina trees in the Beachwood Mangroves Nature Reserve, Durban,South Africa. Data sets were obtained over 5–7 days of relatively dry conditions and over two periods of 5 days during which the swamp was continuously inundated with dilute seawater (< 150 mol m−3NaCl). Gas exchange responses were strongly influenced by photosynthetic photon flux density (PPFD), leaf temperature and leaf to air vapour pressure deficit (Δw). Carbon dioxide exchange was saturated at a PPFD of about 800 μmol m−2 s−1. Maximal CO2 exchange rates ranged from 8.5 to 9.9 μmol m−2 s−1 with no differences between drained and waterlogged conditions. Under drained conditions, leaf conductance,transpiration and internal CO2 concentrations were generally lower, and water use efficiencies higher, than during waterlogging. Continuous waterlogging for 5 days had no adverse effect on CO2 exchange. Xylem water potentials ranged from −1.32to −3.53 MPa during drained and from −1.02 to −2.65 Mpa during waterlogged conditions. These results are discussed in relation to anatomical and metabolic adaptations of A. marina to waterlogging stress. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

13.
Nianzhi Jiao  I-Hsun Ni 《Hydrobiologia》1997,352(1-3):219-230
Geographic and vertical variations of size-fractionated (0.2–1μm, 1–10 μm, and >10 μm) Chlorophyll a (Chl.a) concentration, cyanobacteria abundance and heterotrophic bacteria abundance were investigated at 13 stations from 4°S, 160°W to 30°N, 140°E in November 1993. The results indicated a geographic distribution pattern of these parameters with instances of high values occurring in the equatorial region and offshore areas, and with instance of low values occurring in the oligotrophic regions where nutrients were almost undetectable. Cyanobacteria showed the highest geographic variation(ranging from 27×103 to 16,582×103cell l-1), followed by Chl.a (ranging from 0.048 to 0.178μg l-1), and heterotrophic bacteria (ranging from2.84×103 to 6.50 ×105 cell l-1). Positive correlations were observed between nutrients and Chl.a abundance. Correspondences of cyanobacteria and heterotrophic bacteria abundances to nutrients were less significant than that of Chl.a. The total Chl.a was accounted for 1.0–30.9%, 35.9–53.7%, and 28.1–57.3% by the >10μm, 1–10 μm and 0.2–1 μm fractions respectively. Correlation between size-fractionated Chl.a and nutrients suggest that the larger the cell size, the more nutrient-dependent growth and production of the organism. The ratio of pheophytin to chlorophyll implys that more than half of the >10 μm and about one third of the 1–10 μm pigment-containing particles in the oligotrophic region were non-living fragments, while most of the 1–10 μm fraction was living cells. In the depth profiles, cyanobacteria were distributed mainly in the surface layer, whereas heterotrophic bacteria were abundant from surface to below the euphotic zone. Chl.a peaked at the surface layer (0–20 m) in the equatorial area and at the nitracline (75–100 m) in the oligotrophic regions. Cyanobacteria were not the principle component of the picoplankton. The carbon biomass ratio of heterotroph to phytoplankton was greater than 1 in the eutrophic area and lower than 1 in oligotrophic waters. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

14.
Li H B  Xiao T  Ding Tao  Lü R H 《农业工程》2006,26(4):1012-1019
Four surveys were carried out in August 2001, and in August, September and October 2002 in the southern Yellow Sea, China. Bacterioplankton abundance and biomass were quantified along the transects from Qingdao, China to the Korean island of Jeju. In this paper, the author describes the distribution of bacterioplankton and predatory pressures, and also discusses the impact of Yellow Sea Cold Water Mass (YSCWM) on the distribution of bacterioplankton. The range of Synechococcus biomass was 0.78–33.49 mg C/m3 (on average 6.26 mg C/m3), and heterotrophic bacteria biomass was 1.58–21.25 mg C/m3 (on average 5.79 mg C/m3, n=197). In the vertical direction, the distribution order of Synechococcus biomass was mid-layer>surface>bottom, and the one of heterotrophic bacteria biomass was surface>mid-layer>bottom. The contribution of Synechococcus to the total phytoplankton biomass (SB/PB) ranged from 2% to 99% (on average 42.5%), and the ratio of heterotrophic bacterial biomass to phytoplankton biomass (HBB/PB) was from 5% to 637% (on average 85%). Daily variations of bacterioplankton biomass at anchor stations showed that the maximal Synechococcus biomass was 8.8 times the minimum, and it was 2.8 times in heterotrophic bacterial biomass. The distribution of bacterioplankton accorded with temperature and salinity, and the minimum of bacterioplankton biomass occurred in the YSCWM waters. At the same time, the microzooplankton (20–200 μm) preyed on Synechococcus in these waters, with the ingestion rate of about 0.20–0.42/d.  相似文献   

15.
It has been demonstrated that during the whole year the stems are photosyntheticaly active and capable of assimilating atmospheric CO2. The intensity of photosynthesis varies. During the vegetation period the registered net photosynthesis lasted up to 13 hours per day, and in the leafless period for 2–3 hours a day. Photosynthesis was registered also at temperatures below zero (−3 °C) as a reduced CO2 evolution in light in comparison with darkness. The maximal net photosynthesis values during the vegetation period amounted to 6 up 8 μmol (CO2)·m−2·s−1, and in the leafless period 0.5 – 1 μmol (CO2)·m−2·s−1, and they were close to being up to twice as big as the values obtained of darkness respiration. An increase of the photosynthetic activity of stems preceded the spring development of the leaves.  相似文献   

16.
Annual gross primary productivity in mesotrophic Shahidullah Hall pond (Dhaka, Bangladesh) was 1383.35 g C m−2 y−1 (arithmetic mean). Daily primary productivity (between 1.6 and 6.8 g C m−2 d−1 was correlated with chlorophylla, day length and dissolved silica. Chlorophylla related significantly withk, incident light, SRP, alkalinity and conductivity. A negative correlation existed between biomass and rainfall. Productivity, biomass, conductivity, alkalinity, and SRP increased after mid-winter.k, I k andZ eu varied according seasonally.P max related directly with temperature. Seasonal variation of ∝ B was 0.0049–0.0258 mg C (mg chla mmol PAR)−1 m−2. Q10 was 2.12, community respiration 1334.99 g C m−2 y−1, and the underwater light climate 186.43μE m−2 s−1.  相似文献   

17.
Kotut  Kiplagat  Krienitz  Lothar  Muthuri  Francis M. 《Hydrobiologia》1998,368(1-3):41-59
Temporal changes in phytoplankton chlorophyll a, composition, diversity, biomass (density and fresh weight) and primary production were investigated at the Turkwel Gorge Reservoir (Kenya) over a two year period (1994 and 1995). The phytoplankton properties investigated revealed a seasonal pattern that was very distinct in 1994 and muted in 1995. The wet season was characterized by higher levels of chlorophyll a, biomass and primary production and a lower diversity. A prominent seasonality in 1994 was found to be the result of a higher river inflow volume as compared to 1995. Chlorophyll a changes showed some positive correlation to changes in total nitrogen and total phosphorus. Diversity changes were inversely correlated to changes in total counts (R = −0.84 and −0.96 for 1994 and 1995 respectively). Individual species density changes varied from a distinct seasonal pattern to a nearly uniform density. While the diatom Achnanthes dominated the wet season in 1994, coccoid blue green algae were dominant during most of 1995. Throughout the study period, most biomass was due to the diatoms but with a lower percentage of total biomass in 1995 (40%) as compared to 1994 (88%). The wet season biomass in each year was dominated by the diatoms. Dominance of the intervening period changed irregularly between diatoms, dinoflagellates, green algae and blue green algae. The range of variation in chlorophyll a, total biomass and primary production were; 4.9 to 36.8 μg l-1, 440.14 to 11172.70 mg m-3 and 1.85 to 9.67 g O2 m-2 d-1 in 1994 and 4.9 to 11.5 μg l-1, 486.46 to 1351.39 mg l-1 and 3.08 to 5.41 g O2 m-2 d-1 in 1995 in the same order. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

18.
Photosynthetic rates of green leaves (GL) and green flower petals (GFP) of the CAM plant Dendrobium cv. Burana Jade and their sensitivities to different growth irradiances were studied in shade-grown plants over a period of 4 weeks. Maximal photosynthetic O2 evolution rates and CAM acidities [dawn/dusk fluctuations in titratable acidity] were higher in leaves exposed to intermediate sunlight [a maximal photosynthetic photon flux density (PPFD) of 500–600 μmol m−2 s−1] than in leaves grown under full sunlight (a maximal PPFD of 1 000–1 200 μmol m−2 s−1) and shade (a maximal PPFD of 200–250 μmol m−2 s−1). However, these two parameters of GFP were highest in plants grown under the shade and lowest in full sun-grown plants. Both GL and GFP of plants exposed to full sunlight had lower predawn Fv/Fm [dark adapted ratio of variable to maximal fluorescence (the maximal photosystem 2 yield without actinic irradiation)] than those of shade-grown plants. When exposed to intermediate sunlight, however, there were no significant changes in predawn Fv/Fm in GL whereas a significant decrease in predawn Fv/Fm was found in GFP of the same plant. GFP exposed to full sunlight exhibited a greater decrease in predawn Fv/Fm compared to those exposed to intermediate sunlight. The patterns of changes in total chlorophyll (Chl) content of GL and GFP were similar to those of Fv/Fm. Although midday Fv/Fm fluctuated with prevailing irradiance, changes of midday Fv/Fm after exposure to different growth irradiances were similar to those of predawn Fv/Fm in both GL and GFP. The decreases in predawn and midday Fv/Fm were much more pronounced in GFP than in GL under full sunlight, indicating greater sensitivity in GFP to high irradiance (HI). In the laboratory, electron transport rate and photochemical and non-photochemical quenching of Chl fluorescence were also determined under different irradiances. All results indicated that GFP are more susceptible to HI than GL. Although the GFP of Dendrobium cv. Burana Jade require a lower amount of radiant energy for photosynthesis and this plant is usually grown in the shade, is not necessarily a shade plant.  相似文献   

19.
In order to determine how environmental and physiological factors affect leaf gas exchange in a 9-year-old clonal eucalypt plantation (Eucalyptus grandis Hill ex. Maiden hybrids) in the State of Espirito Santo, Brazil, the diurnal patterns of predawn leaf water potential (Ψpd), and leaf gas exchange were monitored from November 1995 to August 1996. Soil water content (Θ) and microclimatic variables were also recorded. Most of the rainfall during the experimental period occurred from October to December 1995 and from March to April 1996, causing a significant variation in Θ and Ψpd. A high positive correlation (r 2=0.92) was observed between Ψpd and Θ measured at 0.3 m depth from the soil surface. During conditions of high soil water availability, the maximum values of stomatal conductance for water vapor (g s) and net photosynthetic rate (A) were over 0.4 mol m–2 s–2 and l5 μmol m–2 s–1, respectively. The results showed that Ψpd and leaf gas exchange of the examined trees were susceptible to changes in the water content of the upper soil layers, where the major concentration of active roots occur. Multiple linear regression analysis indicated that photosynthetic active radiation (Q), vapor pressure deficit (VPD), atmospheric CO2 molar fraction (C a), and Ψpd were the most important factors controlling g s whereas Q and VPD were the main microclimatic variables controlling A. Received: 5 November 1998 / Accepted: 10 November 1999  相似文献   

20.
The kinetic parameters Km and Vmax for urea uptake by Melosira italica were determined at 160 μeinsteins m−2 s−1 and in the dark. The transport systems showed an affinity for the substrate and a storing capacity in the dark (Km = 65.07 μM; Vmax = 2.18 nmoles 105 cells −1 h−1) greater than under 160 μE m−2 s −1 (Km = 111.2 μM; Vmax = 1.11 nmoles 105 cells−1 h−1). Similarly, a reduction in consumption rate of urea under increasing photon flux densities was observed. The use of an inhibitor (potassium cyanide) indicated that the uptake process requires metabolic energy. That urea transport is more important in darkness, may constitute a survival strategy in which this compound is utilized by cells mainly during heterotrophic growth.  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