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.
Nitzschia seriata Cleve, a common member of marine bottom ice communities in the Arctic, was grown in unialgal batch cultures to test for compensatory mechanisms for the low temperatures (?1.8° C) typical of its natural habitat. The upper lethal limit for growth was between 12° and 15°C, and the optimum was between 6° and 12° C. The Arrhenius function adequately (R2= 73%) fitted the relationship between growth rate and temperature from – 1.6° up to 10° C, with an average Q10 of 1.9 over the entire range. Light-saturated and light-limited rates of photosynthesis (normalized to chlorophyll a or cell carbon) showed complete compensation from 12° to 4° C. Photosynthetic rates, especially at light saturation, declined rapidly at temperatures below 4° C. Susceptibility to photoinhibition was greatest at the lowest growth temperatures. Cellular composition (chlorophyll a, protein, polysaccharide, and lipid contents) was not systematically related to temperature in any simple way, although cell size (carbon per cell) was maximal at the lowest growth temperature. Dark respiration was unmeasurably low (<0.015 day?1) at all growth temperatures. The strategy of adaptation in N. seriata may be characterized as optimizing efficiency and compensation, rather than maximization, of growth rate.  相似文献   

2.
The rate and final germination of four muskmelon cultivars (Cucumis melo) were examined in response to incubation temperatures of 20, 26 and 32°C. Germination was also characterised at 26°C pr 32°C over a range of water potentials from 0 to - 1000 kPa achieved with solutions of polyethylene glycol. The germination of one cultivar, TAM-Uvalde, was consistently slower at 20°C than at 26°C or 32°C. The other three cultivars, Perlita, TAM-Dew and Greenflesh, were inhibited by incubation at 32°C. However, the germination responses of cvs Perlita, TAM-Dew and Greenflesh at 26°C or 32°C improved as water potentials were reduced from 0 to – 200 or – 400 kPa. Cv. TAM-Uvalde was extremely sensitive to water stress and failed to germinate at water potentials below – 600 kPa when incubated at 26°C. The inhibition of germination at low water potentials was partially reversed in all cultivars by increasing the incubation temperature from 26°C to 32°C. It is suggested that the inhibition of germination at 0 kPa (distilled water) was due to a seed coat-mediated barrier to oxygen that could be reversed by removal of the seed coat or exposure to an oxygen-enriched atmosphere.  相似文献   

3.
1. Growth rates of seven species of planktonic algae were determined in culture over a range of temperature from 2 to 35 °C. Additional observations on growth and viability were made for 13 species in the temperature range 20–35 °C. 2. There was a wide range of growth rates between species at their optimal temperatures, from 1.7 divisions day?1 (Asterionella formosa) to 0.3 divisions day?1 (Ceratium furcoides). 3. There were considerable differences between species for growth at low and high temperature. Certain algae, including the diatom A. formosa and the flagellates Cryptomonas marssonii, Dinobryon divergens and Eudorina unicocca var. unicocca, had growth rates of 0.4 divisions day?1 or more at 5 °C. The cyanophyte Tychonema (formerly Oscillatoria) bourrellyi, the xanthophyte Tribonema sp., the desmid Staurastrum cingulum and the large dinoflagellate C. furcoides grew poorly or not at all at this temperature. All 21 species tested could grow at 25 °C, but many – including most of the diatoms, some cyanophytes, and all the flagellates – failed to grow persistently at 30 °C. Only Aphanizomenon flosaquae survived with moderate increase at 35 °C, a lethal temperature for the other species. 4. The applicability was considered of proposed quantitative formulations of the rate‐temperature relationship. Simple exponential relationships applied only to very limited lower ranges of temperature. The relationship proposed by B?lehrádek was a better fit over a wider temperature range, but still excluded rate‐decline at high temperature. 5. The interspecific differences found are of potential significance for restrictions in natural distributions associated with season, altitude (especially above 500 m) and latitude.  相似文献   

4.
Environmental temperature is one of the critical factors affecting fish development. The aim of this study was to examine the impact of three different rearing temperatures (16, 19 and 22°C) throughout the endogenous feeding phase of the Siberian sturgeon Acipenser baerii. This was performed by assessing (a) larval survival and growth; (b) immunofluorescence localization and expression of genes involved in muscle development and growth – myog and Igf1; and (c) stress status through the expression of thermal stress genes – Hsp70, Hsp90α and Hsp90β – and whole body cortisol. Overall survival rate and larval weight did not differ significantly across temperatures. Larvae subjected to 22°C showed faster absorption of the yolk-sac than larvae subjected to 19 or 16°C. Both at schooling and at the end of the trial, larvae reared at 16°C showed significantly lower levels of cortisol than those reared at 19 or 22°C. IGF-1 immunopositivity was particularly evident in red muscle at schooling stage in all temperatures. The expression of all Hsps as well as the myog and Igf1 genes was statistically higher in larvae reared at 16°C but limited to the schooling stage. Cortisol levels were higher in larvae at 22°C, probably because of the higher metabolism demand rather than a stress response. The observed apparent incongruity between Hsps gene expression and cortisol levels could be due to the lack of a mature system. Further studies are necessary, especially regarding the exogenous feeding phase, in order to better understand if this species is actually sensitive to thermal stress.  相似文献   

5.
The objective of this study was to compare the photosynthetic changes during cold acclimation in various plant types able to acquire different degrees of freezing tolerance. Four herbaceous and six woody plants were hardened under natural or artificial conditions and – after determination of their frost resistance (LT50) – the net photosynthetic rate at an ambient CO2 of 33 Pa (Pn33), the dependencies of Pn to light and to CO2 and the room temperature chlorophyll a fluorescence were recorded under optimal conditions. Herbaceous plants acquired freezing tolerances to temperatures between ?10 and ?15°C when hardened at temperatures around 0°C. Most leaves fully developed prior to frost hardening exhibited typical symptoms of senescence after frost hardening. In non-senescing leaves Pn33 was reduced by 15 to 50% mainly due to a reduced stomatal conductance. After hardening at temperatures around ?10°C Brassica survived down to ?24°C, but Pn33 was almost abolished as a result of disturbances in the chloroplasts. After transferring the plants to 20/15°C Pn33 recovered completely within a few days. Woody plants hardened at temperatures around 0°C tolerated – 15 to ?36°C: Pn33 was reduced by 25 to 60% and hardly recovered at 20/15°C. Hardening at ?10°C induced a tolerance of ?32 to n33 was almost totally blocked, but at 20/15°C it returned to the values of the plants hardened at 0°C within a few days. In woody plants disturbances were invariably localized in the chloroplasts. Thus, conifers, and especially Pinus cembra, can survive much lower temperatures than herbaceous plants and, at the same level of freezing tolerance, exhibit appreciably less restriction in relative Pn33.  相似文献   

6.
Tail regeneration was followed for 60 days in 1470 normal (NL), blinded (BL) and pinealectomized (PX) gekkonid lizards, Hemidactylus flaviviridis, fed on cockroaches ad libitum and exposed to continuous light of high intensity (2500 lux), continuous light of low intensity (638 lux), 12 h of light (high intensity) and 12 h of darkness and continuous darkness in summer (March – May; cage temperature 30°C) monsoon (August – October; cage temperature 26°C) and winter (November – January; cage temperature 17°C) seasons. A comparative assessment of the new growth (regenerate) showed that the initiation of regeneration, the daily growth, the total length of tail replaced at the end of regeneration and the total percentage replacement of the lost (autotomized) tails were all enhanced by continuous light and depressed by continuous darkness. Furthermore, there was no significant difference between any of the parameters in NL and BL Hemidactylus; however, PX lizards generally showed retardation in the regeneration process. Seasonal differences in tail regeneration were noted—the best regenerative performance was obtained during the summer months and the worst during the winter months, with the regenerative performance during the monsoon season in between. Thermal and photoperiodic influences on tail regeneration in lizards are discussed.  相似文献   

7.
Photosynthetic responses to temperature and photosynthetically active radiation (PAR) were investigated on the heteromorphic life history stages (macroscopic and microscopic stages) of an edible Japanese brown alga, Cladosiphon okamuranus from the Ryukyu Islands. Measurements were carried out by using optical dissolved oxygen sensors and a pulse‐amplitude modulated fluorometer. Maximum net photosynthetic rates and other parameters of the Photosynthesis – PAR curves at 28°C were somewhat similar in both life history stages, without characteristic photoinhibition at 1000 μmol photons m?2 s?1. Results of oxygenic gross photosynthesis and dark respiration experiments over a temperature range of 8–40°C revealed similar temperature optima for both stages (29.7°C, macroscopic stage; 30.3°C, microscopic stage), which support their observed occurrences in the habitat during summer. Maximum quantum yields of photosystem II (PSII ) (F v /F m ) were relatively stable at low temperatures with the highest at 15.1°C for the macroscopic stage and at 16.5°C for the microscopic stage; but dropped at higher temperatures especially above 28°C. Continuous exposures (6 h) to 200 and 1000 μmol photons m?2 s?1 at 8, 16, and 28°C revealed greater depressions in effective quantum yields of PSII (Φ PSII ) of the microscopic stage at 8°C, as well as its F v /F m that barely increased after 6 h of dark acclimation. Whereas post‐dark acclimation F v /F m of both stages exposed to low PAR fairly recovered at 28°C, suggesting their photosynthetic tolerance to such high temperature. Under natural conditions, both heteromorphic stages of C. okamuranus may persist throughout the year in this region. Beyond its northern limit of distribution, the microscopic stage of this species may suffer from photodamage, as enhanced by low winter temperatures; hence, its restricted occurrence.  相似文献   

8.
To monitor changes in membrane fluidity in Arabidopsis leaves and thylakoid membranes, we investigated the temperature dependence of a chlorophyll fluorescence parameter, minimum fluorescence (Fo), and calculated the threshold temperature [T(Fo)] at which the rise of the fluorescence level Fo was considered to be started. For the modification of membrane fluidity we took three different approaches: (1) an examination of wild‐type leaves initially cultured at room temperature (22°C), then exposed to either a lower (4°C) or higher (35°C) temperature for 5 days; (2) measurements of the shift in T(Fo) by two mutants deficient in fatty acid desaturase genes – fad7 and fad7fad8 and (3) an evaluation of the performance of wild‐type plants when leaves were infiltrated with chemicals that modify fluidity. When wild‐type plants were grown at 22°C, the T(Fo) was 48.3 ± 0.3°C. Plants that were then transferred to a chamber set at 4 or 35°C showed a shift in their T(Fo) to 42.7 ± 0.9°C or 48.9 ± 0.1°C, respectively. Under low‐temperature acclimation, the decline in this putative transition temperature was significantly less in fad7 and fad7fad8 mutants compared with the wild‐type. In both leaf and thylakoid samples, values for T(Fo) were reduced in samples treated with benzyl alcohol, a membrane fluidizer, whereas T(Fo) rose in samples treated with dimethylsulfoxide, a membrane rigidifier. These results indicate that the heat‐induced rise of chlorophyll fluorescence is strongly correlated with the fluidity of thylakoid membranes.  相似文献   

9.
Cultures of Polymyxa graminis were maintained in roots of barley plants grown in sand at different temperatures using Wisconsin soil temperature tanks. At 17 – 20°C, the minimum time from inoculation with cystosori to the production of zoospores from the inoculated roots was 2 – 3 wk. At 11 – 20°C many zoospores were produced but the incubation period was longer at the lower temperatures. Above 20°C little fungal development occurred. The duration of motility of zoospores ranged from c. 1 h to > 24 h. Bovine serum albumen (BSA) prolonged motility but glycine and glucose had no effect or, at higher concentrations, were toxic. Zoospores were rapidly immobilised by zinc ions in solution at or above 10μg/ml. In some experiments BSA added to the zoospore suspension greatly increased transmission of barley yellow mosaic virus (BaYMV) while glucose, glycine and ovalbumen decreased it. When seedlings were incubated with zoospore suspensions for 24 h at different temperatures, BaYMV transmission was high (> 60%) at 10, 15 and 20°C but there was little at 5 or 25°C. In experiments to determine the time taken for zoospore penetration, seedlings were incubated in suspension for different periods of time and then rinsed in zinc sulphate solution to kill free zoospores. Between 3 and 3·5 h was needed for zoospores to establish infection. Transmission occurred equally to plants of various ages between 3 days and 7·5 wk.  相似文献   

10.
1. We examined the detailed temperature dependence (0–40 °C) of bacterial metabolism associated with fine sediment particles from three Danish lowland streams to test if temperature dependence varied between sites, seasons and quality of organic matter and to evaluate possible consequences of global warming. 2. A modified Arrhenius model with reversible denaturation at high temperatures could account for the temperature dependence of bacterial metabolism and the beginning of saturation above 35 °C and it was superior to the unmodified Arrhenius model. Both models overestimated respiration rates at very low temperatures (<5 °C), whereas Ratkowsky's model – the square root of respiration – provided an excellent linear fit between 0 and 30 °C. 3. There were no indications of differences in temperature dependence among samples dominated by slowly or easily degradable organic substrates. Optimum temperature, apparent minimum temperature, Q10‐values for 0–40 °C and activation energies of bacterial respiration were independent of season, stream site and degradability of organic matter. 4. Q10‐values of bacterial respiration declined significantly with temperature (e.g. 3.31 for 5–15 °C and 1.43 for 25–35 °C) and were independent of site and season. Q10‐values of bacterial production behaved similarly, but were significantly lower than Q10‐values of respiration implying that bacterial growth efficiency declined with temperature. 5. A regional warming scenario for 2071–2100 (IPCC A2) predicted that mean annual temperatures will increase by 3.5 °C in the air and 2.2–4.3 °C in the streams compared with the control scenario for 1961–1990. Temperature is expected to rise more in cool groundwater‐fed forest springs than in open, summer‐warm streams. Mean annual bacterial respiration is estimated to increase by 26–63% and production by 18–41% among streams assuming that established metabolism–temperature relationships and organic substrate availability remain the same. To improve predictions of future ecosystem behaviour, we further require coupled models of temperature, hydrology, organic production and decomposition.  相似文献   

11.
Colletotrichum gloeosporioides (Penz.) causing anthracnose is a potent pathogen of chilli resulting in significant yield loss. The in vitro root colonisation study showed an increase in root bacterial count by 10 × 105 colony forming units/cm root for Bacillus subtilis after 15 days of germination. Population level of the antagonist was stable in talc till the 180th day (30°C – 1.6 × 108; 4°C – 1.9 × 108) and in lignite till the 150th day (30°C – 1.5 × 108; 4°C – 1.3 × 108). Combined application of B. subtilis and carbendazim enhanced all biometric parameters with reduction in disease incidence. Soil, seed, root dip and foliar spray treatment significantly enhanced the growth parameters of chilli in B. subtilis inoculated treatments in comparison to the untreated control. Seed application resulted in highest plant fresh weight (76.84 g) and dry weight (34.17 g) compared to the untreated control (50 g and 21 g fresh and dry weight, respectively). Highest plant height ranging from 70 cm (soil application) to 77 cm (dip treatment) with Bacillus inoculation was comparable with carbendazim treatment (61 cm with soil application) and 78 cm (dip treatment) and significantly higher than the untreated control (58.2 cm with soil and 61 cm with application, respectively). Dip treatment resulted in significant increase in root length with B. subtilis (33 cm) and carbendazim (32.5 cm) in comparison to untreated control (15 cm). Co-inoculation of pathogen with B. subtilis (singly and with carbendazim) not only reduced the disease incidence but also improved all the biometric parameters in comparison to challenge inoculation. Root dip application was effective in promoting growth while seed application was effective in disease control.  相似文献   

12.
Chlamydospores are vital asexual resting cells, which allow most of the Fusarium pathogenic strains to retain their longevity, thus ensuring survival of viable reproductive cells. This study suggested that both abiotic – extreme temperature and growth media, and biotic – antagonistic Bacillus amyloliquefaciens SMCD 518 and mycoparasititic Acremonium strictum SMCD 504 are natural stressors able to shift chlamydospores formation in Fusarium graminearum and F. sporotrichioides under in vitro conditions. In F. sporotrichioides, Minimal Conversion Media (MCM) with mannitol supplement induced high chlamydospore size, and chain abundance at optimal 21°C and extreme 37°C temperatures, respectively. F. graminearum showed low chlamydospore formation on MCM–mannitol, even when exposed to 37°C under prolonged 5 days incubation. Generally, F. sporotrichioides has higher chlamydospore abundance, longer chlamydospore chain, and production rapidity compared to F. graminearum in both abiotic and biotic treatments. However, biocontrol bacteria and mycoparasite posed minimal effects on chlamydospore formation, as compared to abiotic stressors, thus controlling the Fusaria but not triggering them to generate chlamydospores as protection shields.  相似文献   

13.
Laboratory investigations were made into the effect of cultivar type, prior cold storage, fumigation temperature and methyl bromide concentration, on the efficacy of a fumigation treatment to control Liriomyza trifolii in chrysanthemum cuttings. The tests related to the standard quarantine treatment used in the UK to control Spodoptera littoralis on imported chrysanthemum cuttings: cold storage for 2 days at 1 – 2°C followed by methyl bromide fumigation at 15°C with a concentration time product (CTP) of 54 g h/m3. L. trifolii larvae, within detached leaves, and 1 – 2 and 2 – 3 day old pupae, were treated. Methyl bromide concentrations of 6·75 or 13·5 g/m3 were used to achieve a range of CTPs and thus obtain accurate dose-response lines and estimates of the LD99 and LD99·9 for each insect stage. Fumigation temperatures were 8, 11 or 15°C. Efficacy of the standard treatments differed between the three cultivars tested, but the LD99 for larvae remained below 54 g h/m3. Decreasing fumigation temperature to 11°C or less increased LD99 values for larvae and pupae and substantially increased variability. There is therefore little scope for using fumigation temperatures of less than 15°C for quarantine purposes. Omitting the cold storage treatment prior to fumigation did not significantly affect efficacy of fumigation. Reducing the methyl bromide concentration from 13·5 to 6·75 g/m3 did not significantly affect the LD99 for larvae but significantly reduced LD99s for pupae.  相似文献   

14.
A study was conducted to determine the effects of six constant temperatures (15, 20, 25, 30, 35 and 40°C) on the post‐diapause embryonic development and the hatching time in three grasshopper species –Omocestus haemorrhoidalis (Charp.), Calliptamus abbreviatus Ikonn. and Chorthippus fallax (Zub.) – from the Inner Mongolian steppe. The results indicate that the species differ in the developmental rates, survival curves and cumulative hatching probabilities. The eggs of O. haemorrhoidalis had the fastest developmental rate with a low developmental threshold temperature of 9.9°C and the sum of effective temperature (SET) 211.2 degree‐days (DD). The corresponding values were 10.9°C and 210.6 DD for C. abbreviatus, 10.5°C and 240.2 DD for Ch. fallax respectively. The SET at which 50% of post‐diapause eggs hatched were 252.0 DD for O. haemorrhoidalis, 262.8 DD for C. abbreviatus, and 273.3 DD for Ch. fallax. The predicted maximal hatch ability of O. haemorrhoidalis (91.17%), C. abbreviatus (75.67%) and Ch. fallax (94.07%) occurred at 23.7, 29.0 and 31.3°C, respectively. The thermal death points of each species were reached at 43.3, 45.0 and 48.6°C. The optimal temperature ranges were 12.2–35.2°C for O. haemorrhoidalis, 21.7–36.3°C for C. abbreviatus and 20.9–41.7°C for Ch. fallax respectively. These results suggest that O. haemorrhoidalis adapt to hatch at a lower temperature range, C. abbreviatus adapt to mid‐temperature range, while Ch. fallax adapt to hatch at a higher temperature range. Although the SET of Ch. fallax is more than that of the other two species, it is not sufficient to explain the hatching sequence of the species in springtime. The results also indicate that Ch. fallax and O. haemorrhoidalis have wider adaptive temperature range than C. abbreviatus.  相似文献   

15.
The d. c. conductivity of five samples of NaDNA in dry state was investigated. All specimens show similar behavior, with specific resistivity at 40°C of the order of 1013–1014 ohm-cm and an energy gap of Ec = 2.04–2.25 eV. Based on the Eley's model for conductivity mechanism, the possibility of conservation of an ordered structure in the dry state (temperature range 20–80°C) is discussed. This is ascribed to the interaction of electron systems of individual bases. From the activation energy values it may be estimated that 17–19 π-electrons take part in the interaction. Photoconductivity measurements indicated activation energies Ef of 0.88–0.96 eV in the visible region.  相似文献   

16.
To understand the effects of low temperature stress on Kappaphycus alvarezii and the responses of antioxidant systems and photosystem II (PSII), behaviour in K. alvarezii thalli exposed to low temperatures (20°C, 17°C and 14°C) for 2 hours was evaluated. Compared with the control at 26°C, activities of some antioxidant enzymes including superoxide dismutase (SOD), peroxidase (POD), ascorbate peroxidase (APX) and the level of antioxidant substance (reduced glutathione) increased in K. alvarezii thalli when exposed to lowered temperatures (20°C, 17°C). Hydroxyl free radical (·OH) scavenging activity of K. alvarezii thalli also increased at 20°C and 17°C compared with the control. This indicated that the resistance to low temperature stress in the antioxidant system of K. alvarezii increased at lowered temperatures of 20°C and 17°C. However, at the lowest temperature (14°C), no significant increases of this algal antioxidant were observed. Under low temperature stress, the maximum quantum yield of PSII photochemistry (FV/FM) and PSII actual photochemical efficiency (ΦPSII) decreased in K. alvarezii thalli, suggesting that the photosynthetic capacity declined. Components of the photosynthetic apparatus (such as the oxygen-evolving complex, light absorption antennas, reaction centres, electron acceptor sides and electron donor sides of PSII) were damaged by low temperature stress to varying degrees. In addition, it was found that low temperature stress led to decreases of both D1 protein and Rubisco LSU (Rubisco large subunit) protein levels. This work is a significant contribution towards understanding the basic mechanism involved in the resistance and the adaptation of K. alvarezii to low temperature stress conditions.  相似文献   

17.
The development period, survival rate, longevity and fecundity of two whiteflies, Bemisia tabaci B‐biotype and Trialeurodes vaporariorum (Homoptera: Aleyrodidae) were compared under different temperature laboratory conditions (15°C, 18°C, 21°C and 24°C). Egg development of B. tabaci B‐biotype was significantly longer compared with that of T. vaporariorum at 15°C, 18°C and 24°C. Significantly longer pseudo‐pupae development and lower survival rate were found in B. tabaci B‐biotype at 15°C compared with those at 18°C, 21°C and 24°C. Significantly higher fecundity was found in B. tabaci B‐biotype at 24°C compared with that at 15°C, 18°C and 21°C. However, the fecundity of T. vaporariorum was significantly lower at 24°C relative to that at 15°C, 18°C and 21°C. Significantly shorter 1st instar larval development was found in T. vaporariorum compared with that of B. tabaci at 15°C and 18°C. Significantly longer 2nd instar larval development was found in B. tabaci and T. vaporariorum at 15°C compared with that at 18°C, 21°C and 24°C. However, significantly shorter 3rd instar larval development was found in T. vaporariorum compared with that of B. tabaci at 15°C, 18°C and 24°C. The adaptive divergence of tolerance to relatively low temperature may be an important factor that results in the interspecific differentiation between the seasonal dynamics of these two whiteflies in China.  相似文献   

18.
The present study describes a strain of Gloeocapsa sp. designated as Gacheva 2007/R‐06/1, originally isolated from a geothermal flow located in Rupite, Bulgaria. To evaluate whether this cyanobacterium is locally adapted to hot environment or has the ability to tolerate lower temperatures, its growth, biochemical composition, enzyme isoforms and activity of the main antioxidant enzymes and proteases were characterized under various temperatures and two irradiance levels. The strain was able to grow over the whole temperature range (15–40°C) under two different photon fluence densities – 132 μmol photons m?2 s?1 (unilateral, low light, LL) and 2 × 132 μmol photons m?2 s?1 (bilateral, high light, HL). The best growth occurred at either 34°C and LL or at 36°C and HL, but significant growth inhibition was noted at 15°C and 40°C. Low temperature treatment (15°C) resulted in higher levels of total protein and an increased activity of manganese superoxide dismutase (MnSOD) and glutathione reductase, as compared to optimum growth temperatures. After simultaneous exposure to 15°C and HL, increases in lipid content and activity of iron superoxide dismutase and catalase (CAT) were also observed. Cultivation of cells at 40°C enhanced MnSOD, CAT and peroxidase activities, regardless of irradiance level. Increased total protein content and protease activity at 40°C was only associated with the HL treatment. Overall, these results indicate that Gloeocapsa sp. strain Gacheva 2007/R‐06/1 used different strategies to enable cells to efficiently acclimate and withstand adverse low or high temperatures. This strain obviously tolerates a wide range of temperatures below its natural habitat temperature, and does not seem to be locally adapted to its original thermal regime. It behaved as a thermotolerant rather than a thermophilic cyanobacterium, which suggests its wider distribution in nature.  相似文献   

19.
The present study scrutinised how far temperature would affect the velocity of the insecticidal activity of Bacillus thuringiensis, as the rapidity of pest control achievements is of a great concern. Third instar Spodoptera littoralis larvae were treated with Bt at three concentration levels under five different temperatures (15°C, 20°C, 25°C, 30°C and 35°C). LT50s were evaluated in each case. The LT50 values showed various levels of reductions as temperature and/or Bt concentration increased, indicating that the velocity of mortality (1/LT50) and/or the rapidity of Bt activity was almost temperature dependant. However, relatively high and low reduction percentages in the LT50 values on the elevation of 5°C were obtained at lower and higher temperature ranges, respectively. The temperature coefficient, Q 10 values, determined within narrow ranges (5°C) showed great reductions when temperature increased from 15°C to 20°C at all Bt concentrations. Raising temperature by 5°C above 20°C or 25°C almost caused similar Q 10 values indicating constant increase in the response of Bt activity within 20–30°C temperature range. Q 10 values over 30°C were comparatively very low. This proved that decrease in Q 10 values due to the rise of temperature was dependant on the starting temperature.  相似文献   

20.
Abstract

Six varieties of lentil (Lens culinaris Medik.) – Asha, Subrata, IPL 406, IPL 81, Lv and Sehore – were exposed to temperatures ranging from 30–50°C which resulted in retarded germination and seedling growth at higher temperatures. Tolerance index and membrane stability tests revealed Sehore and Lv to be susceptible to elevated temperatures while IPL 406, IPL 81, Asha and Subrata were tolerant. Catalase, ascorbate peroxidase and superoxide dismutase showed an initial increase before declining at 50°C, while peroxidase and glutathione reductase activities declined at all temperatures. Lipid peroxidation significantly increased in all varieties. In the tolerant varieties, there was an initial decrease in accumulation of H2O2 followed by an increase from 40°C onwards; however, in the susceptible varieties, accumulation was enhanced at all high temperatures. Ascorbate and glutathione also showed initial increase followed by a decline. Total antioxidant activity was at a maximum at 35–40°C in the tolerant varieties and at 30°C in the susceptible ones. Oxidative stress induced by high temperature was ameliorated by treatment with salicylic acid, abscisic acid or CaCl2, of which salicylic acid was the most effective.  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