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 31 毫秒
1.
Summary Photosynthesis-irradiance relationships and the carbon metabolism of different ice algal assemblages collected from Weddell Sea pack ice were investigated during the EPOS 1 cruise. Infiltration- and interstitial assemblages exhibited the photosynthetic characteristics of high-light adapted ice algae with a mean assimilation number of 1.81±0.93 mg C (mg Chl a)–1 h–1. A higher light harvesting efficiency under light limited conditions (alphaB-value), as well as a lower light intensity for light saturation (IK-value) was determined for the interstitial assemblage. An increase in light intensity from 3.5 to 106 mol m–2s–1 resulted in increased synthesis of polymeric carbohydrates (presumably reserve material) in a band assemblage. However, the absolute incorporation of radiolabel into lipid- and amino acid fractions remained essentially constant over this range of photon flux densities. Light-saturated rates of photosynthesis of three infiltration assemblages under hypersaline conditions (approx. 50 and 110%) decreased by 13–55% (controls: approx. 32–34%). The adverse effect of salinity treatment was much less pronounced under hyposaline conditions (approx. 20), where maximal photosynthetic rates were only slightly decreased (-9%) or even stimulated (14–22%). These observations suggest that sea ice microalgae in the ice edge region of the Weddell Sea during spring, being in a metabolically active stage, may have the potential to initiate or contribute to phytoplankton blooms upon release into the water column.Data presented here were collected during the European Polarstern Study (EPOS) sponsored by the European Science Foundation  相似文献   

2.
Tiina Nõges 《Hydrobiologia》1996,338(1-3):91-103
The material for pigment analysis was collected 1–3 times a year from Lake Peipsi-Pihkva in 1983, 1987, 1988, 1991 and 1992–1995. Concentrations of chlorophyll a, b and c (Chla, Chlb, Chlc), pheopigment (Pheo) and adenosine triphosphate (ATP) were measured biweekly in 1985–1986. The mean of all Chla values was 20.2 mg m–1 (median 13.3 mg m–1) indicating the eutrophic state of the lake. Average Chlb, Chlc, Pheo and carotenoid (Car) contents were 3.7 mg m–3, 4.1 mg m–3, 3.0 mg m–3 and 4.8 mg m–3, respectively. The average Chlb/Chla ratio was 22.9%, Chlc/Chla 23.4%, Pheo/Chla 38%, Car/Chla 37% and ATP/Chla 3%, the medians being 14.3, 13.6, 17.5, 39.4 and 1.9%, respectively. The proportion of Chla in phytoplankton biomass was 0.41%, median 0.32%. There were no significant differences in temperature, oxygen concentration, Chla, and ATP between the surface and bottom water; the lake was polymictic during the vegetation period. The Chla concentration had its first peak in May followed by a decrease in June and July. In late summer Chla increased again achieving its seasonal maximum in late autumn. The ATP concentration was the highest during spring and early summer, decreasing drastically in autumn together with the decline of primary production. ATP/Chla was the highest during the clear water period in June and early July, which coincided also with the high proportion of Chla in phytoplankton biomass. The highest Chla occurred in November (average 37.2 mg m–3) when Secchi transparency was the lowest (1.05 m). Concentrations of Chlb, Chlc and carotenoids were the highest in August, that of Pheo in June. Concentrations of Chla and other pigments were the lowest in the northern part of Lake Peipsi (mean 14.7 mg m–3, median 12.5 mg m–3) and the highest in the southern part of Lake Pihkva (mean 47.9 mg m–3, median 16.3 mg m–3). An increase of Chla and decrease of Secchi depth could be noticed in 1983–1988, while in 1988–1994 the tendency was opposite.  相似文献   

3.
Simulations of continuous ethanol or acetonobutylic fermentations in aqueous two-phase systems show that at high substrate feed concentrations it is possible to obtain solvent productivities about 25–40% higher than in conventional systems with cell recycle if the biomass bleed rate is kept about one tenth of the value of D.List of Symbols a Volumetric fraction of dextran rich phase - B h–1 Bleed rate - D h–1 Dilution rate - P kg m–3 Product concentration - PD kg m–3 h–1 Productivity - S kg m–3 Substrate - X kg m–3 Biomass - Partition coefficient  相似文献   

4.
The effects of light intensity, oxygen concentration, and pH on the rates of photosynthesis and net excretion by metalimnetic phytoplankton populations of Little Crooked Lake, Indiana, were studied. Photosynthetic rates increased from 1.42 to 3.14 mg C·mg–1 chlorophylla·hour–1 within a range of light intensities from 65 to 150E·m–2·sec–1, whereas net excretion remained constant at 0.05 mg C·mg–1 chlorophylla·hour–1. Bacteria assimilated approximately 50% of the carbon released by the phytoplankton under these conditions. Excreted carbon (organic compounds either assimilated by bacteria or dissolved in the lake water) was produced by phytoplankton at rates of 0.02–0.15 mg C·mg–1 chlorophylla·hour–1. These rates were 6%–13% of the photosynthetic rates of the phytoplankton. Both total excretion of carbon and bacterial assimilation of excreted carbon increased at high light intensities whereas net excretion remained fairly constant. Elevated oxygen concentrations in samples incubated at 150E· m–2·sec–1 decreased rates of both photosynthesis and net excretion. The photosynthetic rate increased from 3.0 to 5.0 mg C·mg–1 chlorophylla· hour–1 as the pH was raised from 7.5 to 8.8. Net excretion within this range decreased slightly. Calculation of total primary production using a numerical model showed that whereas 225.8 g C·m–2 was photosynthetically fixed between 12 May and 24 August 1982, a maximum of about 9.3 g C·m–2 was released extracellularly.  相似文献   

5.
Two flight parameters (take-off and duration) and respiration level were measured, in two years in summer and early autumn, in dormant Coccinella septempunctata L. (Coleoptera: Coccinellidae) collected while hidden in grass tussocks in hibernation sites (HID) and in beetles collected on plants (PLA). The duration of tethered flight of HID beetles measured in the laboratory in late August and September 1995 (range of geometric means 190–440 s) was slightly longer than the flight of PLA beetles (80–310 s), both being much longer than trivial flight recorded in beetles foraging for prey during the breeding season (35 s). In general, the flight performance had a tendency to increase in September and to decrease in October.The oxygen consumption in HID beetles increased throughout September 1994 from 430 to 780 l g–1 h–1 and throughout October 1995 from 710 to 1060 l g–1 h–1. This increase is ascribed to a concomitant decrease in diapause intensity. A similar increase was observed also in PLA beetles in 1994 and oxygen consumption was always higher than in HID beetles, most probably due to feeding and digestion in PLA beetles.Laboratory feeding of HID beetles on aphids induced maturation of ovaries and increased oxygen uptake (from 680 to 1110 l g–1 h–1). Feeding on honey and pollen left their oxygen uptake unchanged. Effect of feeding on the flight parameters was mostly not significant. In agreement with its less suitable body shape and usually less distant dormancy sites, C. septempunctata was found a less apt flier than long-distance migrating coccinellid species.  相似文献   

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

7.
Rhizobium trifolii, R. leguminosarum, andR. hedysarum, grownex planta under anoxic conditions in a chemically defined medium, evolve N2O from NO3 , NO2 , and (NH4)2NO3. The amount of nitrous oxide formed after 96 hours is about 0.2M×mg–1 cells d.w. Large availability of organic matter enhances the production of N2O from nitrate by free-livingR. trifolii in peat/sand mixtures. Denitrification of the above species andR. meliloti was detected also in planta. Nitrous oxide production increases almost linearly from 10–45M×mg–1 nodules d.w. when nitrogen-fixing plants are exposed to increasing concentrations of nitrate (1–12M).  相似文献   

8.
Gupta  Rani  Saxena  R. K.  Sharmila  P. 《Current microbiology》1994,29(5):287-289
Cell-bound cholinesterase enzyme activity is reported for the first time in the mycelium ofTrichoderma harzianum. This enzyme hydrolyzes both the acetylcholine and the butyryl thiocholine esters. TheK m andV max for choline ester are 0.69 mM and 1.0 nmol acid released min–1 g–1 protein. However, the thiocholine ester has aK m value of 2.2 mM andV max value of 3.33 nmol product formed min.–1 g–1 protein. The enzyme is inhibited by eserine, a true classical cholinesterase inhibitor.  相似文献   

9.
Age and growth of the whiskery shark, Furgaleus macki, from southwestern Australia were examined using vertebral ageing and tag-recapture data. The readability of bands on the vertebral centra varied markedly between individuals. Four readers were used to make band counts, with the most experienced reader having the lowest index of average percent error and the highest level of agreement with final counts. Marginal increment analysis indicated that opaque bands form in January. With parturition occurring from August to October, size data suggests that the first band is probably formed 15–17 months after birth. The age at maturity was estimated to be 4.5 years for males, and 6.5 years for females. The oldest male was 10.5 years, and oldest female was 11.5 years. Von Bertalanffy growth parameters for males were L =121.5cm fork length, K=0.423 year–1, t 0=–0.472 years, were L =120.7cm fork length, K=0.369 year–1, t 0=–0.544 years for females, and were L =118.1cm fork length, K=0.420 year–1, t 0=–0.491 years for combined sexes. Data from a tag recapture study were analysed using a maximum likelihood method to verify the estimates of growth parameters from vertebral ageing. Von Bertalanffy growth parameters from the tag recapture study were L =128.2cm fork length, K=0.288 year–1, t 0=–0.654 years. The two methods of estimating growth parameters produced similar results, with rapid growth until approximately 5 years of age, after which there was little increase in length.  相似文献   

10.
A field experiment was carried out at two sites off Yucatan State, Southeast Mexico, in order to determine the feasibility of culturing the red seaweedEucheuma isiforme; this was done during May–September 1989. At both sites (Uaymitun and Dzilam) the 25 days harvest and 14 algae per line plant density growth rates (2.21% day–1 and 1.21% day–1, respectively) were significantly higher (p<0.05) than those obtained with other combinations of the two factors tested (50, 75, 100 and 125 days harvest and 9 and 14 algae per line plant density). The mean carrageenan content of the cultured algae was 35.8% and 31.4% at Uaymitun and Dzilam, respectively.  相似文献   

11.
Chikin  S. M.  Tarasova  N. A.  Saralov  A. I.  Bannikova  O. M. 《Microbiology》2003,72(2):213-220
The total population density and the biomass of bacterioplankton, mesozooplankton, and phosphate-accumulating bacteria (PAB) were estimated during the 2000–2001 summer–autumn seasons in the coastal waters of the White and Barents Seas, which are subject to the action of tidal and sea currents, the inflow of riverine waters, and anthropogenic impact. In the shallow estuarine waters with salinities of 6.5–32 near the Chernaya, Pesha, and Pechora River mouths, the population of PAB fluctuated from 0.1 to 9.1 million cells/ml (0–36% of the total bacterial population). In pelagic seawaters, which are low in phosphates (12–50 g/l) and are characterized by an increased iron/phosphorus ratio (2.0–3.6), bacterioplankton amounted to 0.1–1.6 million cells/ml and was mainly represented by small organisms with a volume of 0.08–0.15 m3, commonly lacking intracellular polyphosphates. In the pelagic zone of the Barents Sea, the biomass of mesozooplankton (B z) was comparable with that of bacterioplankton (B b = 39–175 mg/m3), the B b/B z ratio being 1.4–4.6. Off the Varandeiskii, Pechora, and Kolguyev oil terminals, B b increased to 155–300 mg/m3 and the B b/B z ratio rose to 1.4 to 50.3 (with an average value of 20.9), presumably due to the severe anthropogenic impact on these waters. In this case, the dense population of bacterioplankton (0.9–7.6 million cells/ml) was mainly represented by large cells (0.12–0.76 m3 in volume), most of which (3–43% of the total bacterioplankton population) contained polyphosphates. The chemical composition of these waters was characterized by an elevated content of the total phosphorus (65–128 g/l) and by a low iron/phosphorus ratio (0.9–1.2).  相似文献   

12.
Summary Two new forms of the plasma membrane ATP-ase ofMicrococcus lysodeikticus NCTC 2665 were isolated from a sub-strain of the microorganism by polyacrylamide gel electrophoresis. One of them had a mol.wt of 368,000 and a very low specific activity (0.80 µ mol.min–1.mg protein–1) that could not be stimulated by trypsin. This form has been called BI (strain B, inactive). If the electrophoresis was carried out in the presence of reducing agents (i.e., dithiothreitol) and the pH of the effluent maintained at a value of 8.5 another form of the enzyme was obtained. This had a mol.wt of 385,000 and a specific activity of 2.5–5.0 µ mol.min–1.mg protein–1 that could be stimulated by trypsin to 5–10 µ mol.min–1.mg protein–1. This preparation of the ATPase has been called form BA (strain B, enzyme active). The subunit composition of both forms has been studied by sodium dodecyl sulphate and urea gel electrophoresis and compared to that of the enzyme previously purified from the original strain (form A). The three forms of the enzyme had similar and subunits, with mol.wt of about 50,000 and 30,000 dalton, respectively. They also had in common the component(s) of relative mobility 1.0, whose status as true subunit(s) of the enzyme remains yet to be established. However, subunit, that had a mol.wt of about a 52,500 in form A (Andreu et al. Eur. J. Biochem. (1973) 37, 505–515), had a mol.wt similar to in form BI and about 60,000 in form BA. Furthermore BA usually showed two types of this subunit ( and) and an additional peptide chain () with a mol.wt of about 25,000 dalton. This latter subunit seemed to account for the stimulation by trypsin of form BA.Forms BA could be converted to BI by storage and freezing and thawing. Conventional protease activity could not be detected in any of the purified ATPase forms and addition of protease inhibitors to form BA failed to prevent its conversion to form BI. The low activity form (BI) was more stable than the active forms of the enzyme and also differed in its circular dichroism. These results show thatM. lysodeikticus ATPase can be isolated in several forms. Although these variations may be artifacts caused by the purification procedures, they provide model systems for understanding the structural and functional relationships of the enzyme and for drawing some speculations about its functionin vivo.  相似文献   

13.
Recombinant barley -amylase 1 isozyme was constitutively secreted by Saccharomyces cerevisiae. The enzyme was purified to homogeneity by ultrafiltration and affinity chromatography. The protein had a correct N-terminal sequence of His-Gln-Val-Leu-Phe-Gln-Gly-Phe-Asn-Trp, indicating that the signal peptide was efficiently processed. The purified -amylase had an enzyme activity of 1.9 mmol maltose/mg protein/min, equivalent to that observed for the native seed enzyme. The k cat/K m was 2.7 × 102 mM–1.s–1, consistent with those of -amylases from plants and other sources.  相似文献   

14.
The combined effects of water activity (aw) and temperature on mycotoxin production by Penicilium commune (cyclopiazonic acid — CPA) and Aspergillus flavus (CPA and aflatoxins — AF) were studied on maize over a 14-day period using a statistical experimental design. Analysis of variance showed a highly significant interaction (P 0.001) between these factors and mycotoxin production. The minimum aw/temperature for CPA production (2264 ng g–1 P. commune, 709 ng g–1 A. flavus) was 0.90 aw/30 °C while greatest production (7678 ng g–1 P. commune, 1876 ng g–1 A. flavus) was produced at 0.98 aw/20 °C. Least AF (411 ng g–1) was produced at 0.90 aw/20 °C and most (3096 ng g–1) at 0.98 aw/30 °C.  相似文献   

15.
Behavioral and physiological responses to hypoxia were examined in three sympatric species of sharks: bonnethead shark Sphyrna tiburo, blacknose shark, Carcharhinus acronotus, and Florida smoothhound shark, Mustelus norrisi, using closed system respirometry. Sharks were exposed to normoxic and three levels of hypoxic conditions. Under normoxic conditions (5.5–6.4mg l–1), shark routine swimming speed averaged 25.5 and 31.0cm s–1 for obligate ram-ventilating S. tiburo and C. acronotus respectively, and 25.0cm s–1 for buccal-ventilating M. norrisi. Routine oxygen consumption averaged about 234.6 mg O2kg–1h–1 for S. tiburo, 437.2mg O2kg–1h–1 for C. acronotus, and 161.4mg O2 kg–1 h–1 for M. norrisi. For ram-ventilating sharks, mouth gape averaged 1.0cm whereas M. norrisi gillbeats averaged 56.0 beats min–1. Swimming speeds, mouth gape, and oxygen consumption rate of S. tiburo and C. acronotus increased to a maximum of 37–39cm s–1, 2.5–3.0cm and 496 and 599mg O2 kg–1 h–1 under hypoxic conditions (2.5–3.4mg l–1), respectively. M. norrisi decreased swimming speeds to 16cm s–1 and oxygen consumption rate remained similar. Results support the hypothesis that obligate ram-ventilating sharks respond to hypoxia by increasing swimming speed and mouth gape while buccal-ventilating smoothhound sharks reduce activity.  相似文献   

16.
Warburg showed in 1929 that the photochemical action spectrum for CO dissociation from cytochrome c oxidase is that of a heme protein. Keilin had shown that cytochrome a does not react with oxygen, so he did not accept Warburg's view until 1939, when he discovered cytochrome a 3. The dinuclear cytochrome a 3-CuB unit was found by EPR in 1967, whereas the dinuclear nature of the CuA site was not universally accepted until oxidase crystal structures were published in 1995. There are negative redox interactions between cytochrome a and the other redox sites in the oxidase, so that the reduction potential of a particular site depends on the redox states of the other sites. Calculated electron-tunneling pathways for internal electron transfer in the oxidase indicate that the coupling-limited rates are 9×105 (Cu A a) and 7×106 s–1 (a a 3); these calculations are in reasonable agreement with experimental rates, after corrections are made for driving force and reorganization energy. The best CuA-a pathway starts from the ligand His204 and not from the bridging sulfur of Cys196, and an efficient a-a 3 path involves the heme ligands His378 and His376 as well as the intervening Phe377 residue. All direct paths from CuA to a 3 are poor, indicating that direct CuA a 3 electron transfer is much slower than the CuA a reaction. The pathways model suggests a means for gating the electron flow in redox-linked proton pumps.  相似文献   

17.
In a recent study on the degradation of N,N-dibutylurea (DBU), a breakdown product of benomyl [methyl 1-(butylcarbamoyl)-2-benzimidazole carbamate], the active ingredient in Benlate® fungicides, degradation half-lives of 1.4–46.5days were observed across several soils incubated at various combinations of soil moisture potential (–0.03 and –0.1MPa) and temperature (23, 33, and 44°C) for a single DBU application of 0.08 and 0.8 g g–1 (Lee et al. 2004). However, Benlate® can be applied as often as every 7days resulting in the repeated application of DBU likely to be present in the Benlate® over a growing season. In this study, the effect of seven repeated DBU applications on mineralization rate was investigated in two soils, which encompass the range in rates previously observed. For the slower degrading soil, repeated DBU application increased mineralization from 0.029 to 0.99day–1 at the 0.08 g g–1 rate, and 0.037 to 0.89day–1 at the 0.8 g g–1 rate. For the faster degrading soil, effects on mineralization of repeated DBU applications were small to negligible. For the latter soil, the effect on mineralization of applied DBU concentrations from 0.0008 to 80 g g–1 was also investigated. Mineralization rates decreased from 0.43 to 0.019day–1 with increasing DBU concentrations. However, the amount of DBU mineralized by day 70 was similar across concentrations and averaged 83% of applied. Microbial respiration was not affected by increasing DBU concentrations. These findings support the supposition that DBU is readily degraded by soil microorganisms, thus unlikely to accumulate in agricultural soils.  相似文献   

18.
The seasonal variation in primary production, individual numbers, and biomass of phyto- and zooplankton was studied in the River Danube in 1981. The secondary production of two dominant zooplankton species (Bosmina longirostris and Acanthocyclops robustus) was also estimated. In the growing season (April–Sept.) individual numbers dry weights and chlorophyll a contents of phytoplankton ranged between 30–90 × 106 individuals, l–1, 3–12 mg l–1, and 50–170 µg l–1, respectively. Species of Thalassiosiraceae (Bacillariophyta) dominated in the phytoplankton with a subdominance of Chlorococcales in summer. Individual numbers and dry weights of crustacean zooplankton ranged between 1400–6500 individuals m–3, and 1.2–12 mg m–3, respectively. The daily mean gross primary production was 970 mg C m–3 d–1, and the net production was 660 mg C m–3 d–1. Acanthocyclops robustus populations produced 0.2 mg C m–3 d–1 as an average, and Bosmina longirostris populations 0.07 mg C m–3 d–1. The ecological efficiency between phytoplankton and crustacean zooplankton was 0.03%.  相似文献   

19.
D. H. Greer  W. A. Laing 《Planta》1992,186(3):418-425
Kiwifruit (Actinidia deliciosa (A. Chev.) C.F. Liang et A.R. Ferguson) plants grown in an outdoor enclosure were exposed to the natural conditions of temperature and photon flux density (PFD) over the growing season (October to May). Temperatures ranged from 14 to 21° C while the mean monthly maximum PFD varied from 1000 to 1700 mol · m–2 · s–1, although the peak PFDs exceeded 2100 mol · m–2 · s–1. At intervals, the daily variation in chlorophyll fluorescence at 692 nm and 77K and the photon yield of O2 evolution in attached leaves was monitored. Similarly, the susceptibility of intact leaves to a standard photoinhibitory treatment of 20° C and a PFD of 2000 mol · m–2 · s–1 and the ability to recover at 25° C and 20 mol · m–2 · s–2 was followed through the season. On a few occasions, plants were transferred either to or from a shade enclosure to assess the suceptibility to natural photoinhibition and the capacity for recovery. There were minor though significant changes in early-morning fluorescence emission and photon yield throughout the growing season. The initial fluorescence, Fo, and the maximum fluorescence, Fm, were, however, significantly and persistently different from that in shade-grown kiwifruit leaves, indicative of chronic photoinhibition occurring in the sun leaves. In spring and autumn, kiwifruit leaves were photoinhibited through the day whereas in summer, when the PFDs were highest, no photoinhibition occurred. However, there was apparently no non-radiative energy dissipation occurring then also, indicating that the kiwifruit leaves appeared to fully utilize the available excitation energy. Nevertheless, the propensity for kiwifruit leaves to be susceptible to photoinhibition remained high throughout the season. The cause of a discrepancy between the severe photoinhibition under controlled conditions and the lack of photoinhibition under comparable, natural conditions remains uncertain. Recovery from photoinhibition, by contrast, varied over the season and was maximal in summer and declined markedly in autumn. Transfer of shade-grown plants to full sun had a catastrophic effect on the fluorescence characteristics of the leaf and photon yield. Within 3 d the variable fluorescence, Fv, and the photon yield were reduced by 80 and 40%, respectively, and this effect persisted for at least 20 d. The restoration of fluorescence characteristics on transfer of sun leaves to shade, however, was very slow and not complete within 15 d.Abbreviations and Symbols Fo, Fm, Fv initial, maximum, variable fluorescence - Fi Fv at t = 0 - F Fv at t = - PFD photon flux density - PSII photosystem II - leaf absorptance ratio - (a photon yield of O2 evolution (absorbed basis) - i a at t = 0 - a at t = We thank Miss Linda Muir and Amanda Yeates for their technical assistance in this study.  相似文献   

20.
Summary The kinetics ofBordetella pertussis growth was studied in a glutamate-limited continuous culture. Growth kinetics corresponded to Monod's model. The saturation constant and maximum specific growth rate were estimated as well as the energetic parameters, theoretical yield of cells and maintenance coefficient. Release of pertussis toxin (PT) and lipopolysaccharide (LPS) were growth-associated. In addition, they showed a linear relationship between them. Growth rate affected neither outer membrane proteins nor the cell-bound LPS pattern.Nomenclature X cell concentration (g L–1) - specific growth rate (h–1) - m maximum specific growth rate (h–1) - D dilution rate (h–1) - S concentration of growth rate-limiting nutrient (glutamate) (mmol L–1 or g L–1) - Ks substrate saturation constant (mol L–1) - ms maintenance coefficient (g g–1 h–1) - Yx/s theoretical yield of cells from glutamate (g g–1) - Yx/s yield of cells from glutamate (g g–1) - YPT/s yield of soluble PT from glutamate (mg g–1) - YKDO/s yield of cell-free KDO from glutamate (g g–1) - YPT/x specific yield of soluble PT (mg g–1) - YKDO/x specific yield of cell-free KDO (g g–1) - qPT specific soluble PT production rate (mg g–1 h–1) - qKDO specific cell-free KDO production rate (g g–1 h–1)  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