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.
Consequences of warming and postwarming events on photosynthetic thermotolerance (PT) and photoprotective responses in tropical evergreen species remain elusive. We chose Citrus to answer some of the emerging questions related to tropical evergreen species' PT behaviour including (i) how wide is the genotypic variation in PT? (ii) how does PT respond to short-term warming and (iii) how do photosynthesis and photoprotective functions respond over short-term warming and postwarming events? A study on 21 genotypes revealed significant genotypic differences in PT, though these were not large. We selected five genotypes with divergent PT and simulated warming events: Tmax 26/20°C (day-time highest maximum/night-time lowest maximum) (Week 1) < Tmax 33/30°C (Week 2) < Tmax 36/32°C (Week 3) followed by Tmax 26/16°C (Week 4, recovery). The PT of all genotypes remained unaltered despite strong leaf megathermy (leaf temperature > air temperature) during warming events. Though moderate warming showed genotype-specific stimulation in photosynthesis, higher warming unequivocally led to severe loss in net photosynthesis and induced higher nonphotochemical quenching. Even after a week of postwarming, photoprotective mechanisms strongly persisted. Our study points towards a conservative PT in evergreen citrus genotypes and their need for sustaining higher photoprotection during warming as well as postwarming recovery conditions.  相似文献   

2.
Summary The adsorption behaviour of cellulase fromTrichoderma viride on microcrystalline celluloses with different specific surface areas was studied. The adsorption was found to fit a Langmuir isotherm. There was an increase in the maximum adsorption amount (Amax) as the specific surface area of microcrystalline cellulose increased. The values of Amax and adsorption equilibrium constant (K) decreased with increasing temperature. Thermodynamic parameters in adsorption were calculated from K. It was found from the enthalpy of adsorption, that van der Waals-Type interaction was responsible for adsorption of cellulase on microcrystalline cellulose. The adsorption process was exothermic and an adsorption enthalpy-controlled reaction.  相似文献   

3.
Temperature dependence of two parameters in a photosynthesis model   总被引:7,自引:2,他引:5  
The temperature dependence of the photosynthetic parameters Vcmax, the maximum catalytic rate of the enzyme Rubisco, and Jmax, the maximum electron transport rate, were examined using published datasets. An Arrehenius equation, modified to account for decreases in each parameter at high temperatures, satisfactorily described the temperature response for both parameters. There was remarkable conformity in Vcmax and Jmax between all plants at Tleaf < 25 °C, when each parameter was normalized by their respective values at 25 °C (Vcmax0 and Jmax0), but showed a high degree of variability between and within species at Tleaf > 30 °C. For both normalized Vcmax and Jmax, the maximum fractional error introduced by assuming a common temperature response function is < ± 0·1 for most plants and < ± 0·22 for all plants when Tleaf < 25 °C. Fractional errors are typically < ± 0·45 in the temperature range 25–30 °C, but very large errors occur when a common function is used to estimate the photosynthetic parameters at temperatures > 30 °C. The ratio Jmax/Vcmax varies with temperature, but analysis of the ratio at Tleaf = 25 °C using the fitted mean temperature response functions results in Jmax0/Vcmax0 = 2·00 ± 0·60 (SD, n = 43).  相似文献   

4.
In this study, we have examined several physiological, biochemical and morphological features of Buddleja davidii plants growing at 1300 m above sea level (a.s.l.) and 3400 m a.s.l., respectively, to identify coordinated changes in leaf properties in response to reduced CO2 partial pressure (Pa). Our results confirmed previous findings that foliar δ13C, photosynthetic capacity and foliar N concentration on a leaf area basis increased, whereas stomatal conductance (gs) decreased with elevation. The net CO2 assimilation rate (Amax), maximum rate of electron transport (Jmax) and respiration increased significantly with elevation, although no differences were found in carboxylation efficiency of Rubisco (Vcmax). Consequently, also the Jmax to Vcmax ratio was significantly increased by elevation, indicating that the functional balance between Ribulose‐1,5‐biphosphate (RuBP) consumption and RuBP regeneration changes as elevation increases. Our results also indicated a homeostatic response of CO2 transfer conductance inside the leaf (mesophyll conductance, gm) to increasing elevation. In fact, with elevation, gm also increased compensating for the strong decrease in gs and, thus, in the Pi (intercellular partial pressure of CO2) to Pa ratio, leading to similar chloroplast partial pressure of CO2 (Pc) to Pa ratio at different elevations. Because there were no differences in Vcmax, also A measured at similar PPFD and leaf temperature did not differ statistically with elevation. As a consequence, a clear relationship was found between A and gm, and between A and the sum of gs and gm. These data suggest that the higher dry mass δ13C of leaves at the higher elevation, indicative of lower long‐term Pc/Pa ratio, cannot be attributed to changes either in diffusional resistances or in carboxylation efficiency. We speculate that because temperature significantly decreases as the elevation increases, it dramatically affects CO2 diffusion and hence Pc/Pa and, consequently, is the primary factor influencing 13C discrimination at high elevation.  相似文献   

5.
Water‐use efficiency (WUE) has been recognized as an important characteristic of ecosystem productivity, which links carbon (C) and water cycling. However, little is known about how WUE responds to climate change at different scales. Here, we investigated WUE at leaf, canopy, and ecosystem levels under increased precipitation and warming from 2005 to 2008 in a temperate steppe in Northern China. We measured gross ecosystem productivity (GEP), net ecosystem CO2 exchange (NEE), evapotranspiration (ET), evaporation (E), canopy transpiration (Tc), as well as leaf photosynthesis (Pmax) and transpiration (Tl) of a dominant species to calculate canopy WUE (WUEc=GEP/T), ecosystem WUE (WUEgep=GEP/ET or WUEnee=NEE/ET) and leaf WUE (WUEl=Pmax/Tl). The results showed that increased precipitation stimulated WUEc, WUEgep and WUEnee by 17.1%, 10.2% and 12.6%, respectively, but decreased WUEl by 27.4%. Climate warming reduced canopy and ecosystem WUE over the 4 years but did not affect leaf level WUE. Across the 4 years and the measured plots, canopy and ecosystem WUE linearly increased, but leaf level WUE of the dominant species linearly decreased with increasing precipitation. The differential responses of canopy/ecosystem WUE and leaf WUE to climate change suggest that caution should be taken when upscaling WUE from leaf to larger scales. Our findings will also facilitate mechanistic understanding of the C–water relationships across different organism levels and in projecting the effects of climate warming and shifting precipitation regimes on productivity in arid and semiarid ecosystems.  相似文献   

6.
Summary Parameters of thermal death were determined in 10 strains of yeast species whose maximum temperatures for growth (T max) ranged from 22 to 49°C. Arrhenius plots of the specific thermal death rates (k d) formed a positional sequence at the level of the experimental points that corrresponded in all but one case to the sequence of the respective T max values. Extrapolated k d values at higher or lower temperatures no longer formed this sequence.The correlation of the temperature functions with T max could be characterized in terms of a new activation parameter, for which the name thermal death activation constant is introduced. It has the following form: T.D.A. – S where H and S are respectively the apparent heat and entropy of activation of thermal death and n is the number of degrees above T max (expressed in °K) at which the T.D.A. constant exists.Seven mesophilic yeasts had a T.D.A. constant between 72 and 79 calxmol-1 degree-1 at n values between 1 and 4°. This suggested that the destructive process that limits k d in these strains is of the same species as one that contributes to the establishment of T max. Two psychrophilic yeasts apparently had a similar T.D.A. constant but at a high n value (about 12.5°C) which suggested that in these strains T max is governed by a destructive process unrelated to the one that underlies thermal death. The strain of the nearly thermophilic Hansenula angusta (T max 49°C) did not fit in either group.The significance of the T.D.A. constant is discussed and expressions for H and S in terms of bond activation parameters are proposed.  相似文献   

7.
The effect of 30 days of acclimation at 5°C and of a semiweekly series of short severe cold exposures (Tb 20–30°C) on metabolic capacity (Mmax) was measured in Alaskan meadow voles(Microtus pennsylvanicus tananaensis) and Wisconsin deer mice(Peromyscus maniculatus bairdii). Meadow voles, with an Mmax of 12–14 ml/(g.h) or 8–9 met (Mmax/Mst), showed little response to either treatment. In deer mice, however, acclimation at 5°C increased Mmax by about half (from 11.0 to 15.4 ml/(g.h) or from 6.0 to 9.1 met). In 25°C-acclimated deer mice 7 severe cold exposures produced a similar increase of which about half was seen with the first 2 exposures. In 5°C-acclimated deer mice, Mmax averaged a 0.3 ml/(g.h) increase for each cold exposure to reach a level of 19 ml/(g.h) or 11 met after 6 weeks.  相似文献   

8.
The kinetics of the torque-velocity (T-ω) relationship after aerobic exercise was studied to assess the effect of fatigue on the contractile properties of muscle. A group of 13 subjects exercised until fatigued on a cycle ergometer, at an intensity which corresponded to 60% of their maximal aerobic power for 50 min (MAP60%); ten subjects exercised until fatigued at 80% of their maximal aerobic power for 15 min (MAP80%). Of the subjects 7 exercised at both intensities with at least a 1-week interval between sessions. Pedalling rate was set at 60 rpm. The T-ω relationship was determined from the velocity data collected during all-out sprints against a 19 N · m braking torque on the same ergometer, according to a method proposed previously. Maximal theoretical velocity (ω0) and maximal theoretical torque (T 0) were estimated by extrapolation of the linear T-ω relationship. Maximal power (P max) was calculated from the values of T 0 and ω0 (P max = 0.25 ω0T 0). The T-ω relationships were determined before, immediately after and 5 and 10 min after the aerobic exercise. The kinetics of ω0, T 0 and P max was assumed to express the effects of fatigue on the muscle contractile properties (maximal shortening velocity, maximal muscle strength and maximal power). Immediately after exercise at MAP60% a 7.8% decrease in T 0 and 8.8% decrease in P max was seen while the decrease in ω0 was nonsignificant, which suggested that P max decreased in the main because of a loss in maximal muscle strength. In contrast, MAP80% induced a 8.1% decrease in ω0 and 12.8% decrease in P max while the decrease in T 0 was nonsignificant, which suggested that the main cause of the decrease in P max was probably a slowing of maximal shortening velocity. The short recovery time of the T-ω relationship suggests that the causes of the decrease of torque and velocity are processes which recover rapidly. Accepted: 25 November 1996  相似文献   

9.
Air temperatures of greater than 35 °C are frequently encountered in groundnut‐growing regions, especially in the semi‐arid tropics. Such extreme temperatures are likely to increase in frequency under future predicted climates. High air temperatures result in failure of peg and pod set due to lower pollen viability. The response of pollen germination and pollen tube growth to temperature was quantified in order to identify differences in pollen tolerance to temperature among 21 groundnut genotypes. Plants were grown from sowing to harvest in a poly‐tunnel under an optimum temperature of 28/22 °C (day/night). Pollen was collected at anther dehiscence and was exposed to temperatures from 10° to 47·5 °C at 2·5 °C intervals. The results showed that a modified bilinear model most accurately described the response to temperature of percentage pollen germination and maximum pollen tube length. Genotypes were found to range from most tolerant to most susceptible based on both pollen characters and membrane thermostability. Mean cardinal temperatures (Tmin, Topt and Tmax) averaged over 21 genotypes were 14·1, 30·1 and 43·0 °C for percentage pollen germination and 14·6, 34·4 and 43·4 °C for maximum pollen tube length. The genotypes 55‐437, ICG 1236, TMV 2 and ICGS 11 can be grouped as tolerant to high temperature and genotypes Kadiri 3, ICGV 92116 and ICGV 92118 as susceptible genotypes, based on the cardinal temperatures. The principal component analysis identified maximum percentage pollen germination and pollen tube length of the genotypes, and Tmax for the two processes as the most important pollen parameters in describing a genotypic tolerance to high temperature. The Tmin and Topt for pollen germination and tube growth, rate of pollen tube growth were less predictive in discriminating genotypes for high temperature tolerance. Genotypic differences in heat tolerance‐based on pollen response were poorly related (R2 = 0·334, P = 0·006) to relative injury as determined by membrane thermostability.  相似文献   

10.
Summary Well watered plants of Vigna unguiculata (L.) Walp cv. California Blackeye No. 5 had maximum photosynthetic rates of 16 mol m-2 s-1 (at ambient CO2 concentration and environmental parameters optimal for high CO2 uptake). Leaf conductance declined with increasing water vapour concentration difference between leaf and air (w), but it increased with increasing leaf temperature at a constant small w. When light was varied, CO2 assimilation and leaf conductance were correlated linearly. We tested the hypothesis that g was controlled by photosynthesis via intercellular CO2 concentration (c i). No unique relationship between (1) c i, (2) the difference between ambient CO2 concentration (c a) and c i, namely c a-c i, or (3) the c i/c a ratio and g was found. g and A appeared to respond to environmental factors fairly independently of each other. The effects of different rates of soil drying on leaf gas exchange were studied. At unchanged air humidity, different rates of soil drying were produced by using (a) different soils, (b) different irrigation schemes and (c) different soil volumes per plant. Although the soil dried to wilting point the relative leaf water content was little affected. Different soil drying rates always resulted in the same response of photosynthetic capacity (A max) and corresponding leaf conductance (g(Amax)) when plotted against percent relative plant-extractable soil water content (W e %) but the relationship with relative soil water content (W e ) was less clear. Above a range of W e of 15%–25%, A max and g(Amax) were both high and responded little to decreasing W e . As soon as W e fell below this range, A max and g(Amax) declined. The data suggest root-to-leaf communication not mediated via relative leaf water content. However, g(Amax) was initially more affected than A max.List of abbreviations A CO2 assimilation - A max photosynthetic capacity at favourable ambient conditions - c a CO2 concentration of the air in the leaf chamber - c i intercellular - CO2 concentration - E transpiration - g leaf conductance - g(Amax) leaf conductance corresponding to photosynthetic capacity - I photon flux rate - T l leaf temperature - W e relative plant-extractable soil water content - W e absolute plant-extractable soil water content - W l relative leaf water content - W s relative soil water content - w difference in water vapour mole fraction between leaf and air - leaf water potential  相似文献   

11.
Pigeon flight in a wind tunnel   总被引:2,自引:0,他引:2  
Summary Core temperatureT c, breast temperatureT s–br and leg temperatureT s–1 were measured simultaneously in pigeons during rest and flight in a wind tunnel, using thermistors.MeanT c at rest is 39.8±0.7°C and is independent of ambient temperatureT a (10–30°C). In the first minutes of flight,T c increases to 1.5–3.0°C above resting level and remains at this higher level. This hyperthermia increases withT a (v=const.). It is±constant in the lowT a range (10.6–13.9°C) at flight speeds v ranging from 10–18 m s–1 and normal body mass, but increases with v and elevated body mass in the highT a range (23.7–28.8°C). T s–1 is adapted toT a at rest and increases in flight up to 3–4°C belowT c. This increase inT s–1 is linear toT a. T s–br is always lower thanT c, in extreme cases reaching restingT c in flight.Supported by the Deutsche Forschungsgemeinschaft  相似文献   

12.
Plants differ in how much the response of net photosynthetic rate (P N) to temperature (T) changes with the T during leaf development, and also in the biochemical basis of such changes in response. The amount of photosynthetic acclimation to T and the components of the photosynthetic system involved were compared in Arabidopsis thaliana and Brassica oleracea to determine how well A. thaliana might serve as a model organism to study the process of photosynthetic acclimation to T. Responses of single-leaf gas exchange and chlorophyll fluorescence to CO2 concentration measured over the range of 10–35 °C for both species grown at 15, 21, and 27 °C were used to determine the T dependencies of maximum rates of carboxylation (VCmax), photosynthetic electron transport (Jmax), triose phosphate utilization rate (TPU), and mesophyll conductance to carbon dioxide (gm). In A. thaliana, the optimum T of P N at air concentrations of CO2 was unaffected by this range of growth T, and the T dependencies of VCmax, Jmax, and gm were also unaffected by growth T. There was no evidence of TPU limitation of P N in this species over the range of measurement conditions. In contrast, the optimum T of P N increased with growth T in B. oleracea, and the T dependencies of VCmax, Jmax, and gm, as well as the T at which TPU limited P N all varied significantly with growth T. Thus B. oleracea had much a larger capacity to acclimate photosynthetically to moderate T than did A. thaliana.  相似文献   

13.
Characterization of Clostridium thermocellum JW20   总被引:9,自引:3,他引:6       下载免费PDF全文
Clostridium thermocellum JW20 (ATCC 31549), which was isolated from a Louisiana cotton bale, grew on cellulose, cellobiose, and xylooligomers and, after adaptation, on glucose, fructose, and xylose in the pH range of 7.5 to 6.1 with Topt of 60°C, Tmax of 69°C, and Tmin of above 28°C. Doubling times during growth on cellulose and cellobiose were 6.5 and 2.5 h, respectively. The G+C content of the DNA was 40 mol% (chemical analysis). Growth on cellulose as substrate was totally inhibited in the presence of more than 125 mM sodium sulfate, 300 mM sodium chloride, 250 mM potassium chloride, 200 mM calcium chloride, 125 mM magnesium chloride, 40 mM lactate, or 250 mM acetate. The ratio of the fermentation products ethanol to acetate plus H2 decreased when the culture was agitated. Agitation otherwise increased the rate of cellulose degradation in a growing culture but not under nongrowth conditions or with cell-free culture supernatant containing the extracellular cellulase. Shaking lowered the concentration of H2 in the culture broth and thus minimized inhibition by the H2 formed. Externally added H2 caused an increased formation of ethanol during growth on cellulose or cellobiose. However, at an atmospheric pressure as high as 355 kPa (50 lb/in2), H2 did not cause significant growth inhibition beyond an increasing lag phase (up to 24 h). Several criteria to specifically prove the purity of C. thermocellum cultures were suggested.  相似文献   

14.
15.
Aims: Non‐sigmoid growth curves of Escherichia coli obtained at constant temperatures near the maximum growth temperature (Tmax) were previously explained by the coexistence of two subpopulations, i.e. a stress‐sensitive and a stress‐resistant subpopulation. Mathematical simulations with a heterogeneous model support this hypothesis for static experiments at 45°C. In this article, the behaviour of E. coli, when subjected to a linearly increasing temperature crossing Tmax, is studied. Methods and Results: Subpopulation dynamics are studied by culturing E. coli K12 MG1655 in brain heart infusion broth in a bioreactor. The slowly increasing temperature (°C h?1) starting from 42°C results in growth up to 60°C, a temperature significantly higher than the known Tmax. Given some additional presumptions, mathematical simulations with the heterogeneous model can describe the dynamic experiments rather well. Conclusions: This study further confirms the existence of a stress‐resistant subpopulation and reveals the unexpected growth of E. coli at temperatures significantly higher than Tmax. Significance and Impact of the Study: The growth of the small stress‐resistant subpopulation at unexpectedly high temperatures asks for a revision of currently applied models in food safety and food quality strategies.  相似文献   

16.
This study investigated the question: is core temperature measurement influenced by whether exercise involves predominantly upper- or lower-body musculature? Healthy men were allocated to three groups: treadmill ergometry (T) n=4, cycle ergometry (C) n=6 and arm crank ergometry (AC) n=5. Subjects underwent an incremental exercise test to exhaustion on an exercise-specific ergometer to determine maximum/peak oxygen consumption (O2max). One week later subjects exercised for 36 min on the same ergometer at approximately 65% O2max while temperatures at the rectum (T re) and esophagus (T es) were simultaneously measured. The O2max (l · min−1) for groups T [4.76 (0.50)] and C [4.35 (0.30)] was significantly higher than that for the AC group [2.61 (0.24)]. At rest, T re was significantly higher than T es in all groups (P<0.05). At the end of submaximal exercise in the C group, T re [38.32 (0.11)°C] was significantly higher than T es [38.02 (0.12)°C, P<0.05]. No significant differences between T re and T es at the end of exercise were noted for AC and T groups. The temperature difference (T diff) between T re and T es was dissimilar at rest in the three groups; however, by the end of exercise T diff was approximately 0.2°C for each of the groups, suggesting that at the end of steady-state exercise T re can validly be used to estimate core temperature. Accepted: 3 November 1997  相似文献   

17.
Summary The capacity for sustained, terrestrial locomotion in the cockroach. Blaberus discoidalis, was determined in relation to running speed, metabolic cost, aerobic capacity, and ambient temperature (T a=15, 23, and 34°C; acclimation temperature=24°C). Steady-state thoracic temperature (T tss) increased linearly with speed at each T a.The difference between T tss and T awas similar at each experimental temperature with a maximum increase of 7°C. Steady-state oxygen consumption (VO2ss) increased linearly with speed at each T aand had a low thermal dependence (Q10=1.0-1.4). The minimum cost of locomotion (the slope of the VO2ss versus speed function) was independent of T a.Cockroaches attained a maximal oxygen consumption (VO2max). increased with T afrom 2.1 ml O2·g-1·h-1 at 15°C to 4.9 ml O2·g-1·h-1 at 23°C, but showed no further increase at 34°C, VO2max increased 23-fold over resting VO2 at 23°C, 10-fold at 34°C, and 15-fold at 15°C. Endurance correlated with the speed at which VO2max was attained (MAS, maximal aerobic speed). Temperature affected the kinematics of locomotion. compared to cockroaches running at the same speed, but higher temperatures (23–34°C), low temperature (15°C) increased protraction time, reduced stride frequency, and reduced stability by increasing body pitching. The thermal independence of the minimum cost of locomotion (Cmin), the low thermal dependence of VO2ss (i.e., y-intercept of the VO2ss versus speed function), and a typical Q10 of 2.0 for VO2max combined to increase MAS and endurance in B. discoidalis when T awas increased from 15 to 23°C. Exerciserelated endothermy enabled running cockroaches to attain a greater VO2max, metabolic scope, and endurance capacity at 23°C than would be possible if T tss remained equal to T a. The MAS of B. discoidalis was similar to that of other arthropods that use trachea, but was 2-fold greater than ectotherms, such as salamanders, frogs, and crabs of a comparable body mass.Abbreviations T a ambient temperature - T t thoracic temperature - T tss steady state thoracic temperature during exercise - T trest thoracic temperature during rest - VO2 oxygen consumption - VO2rest oxygen consumption during rest - VO2ss steady-state oxygen consumption during exercise - VO2max maximal oxygen consumption; MAS maximum aerobic speed - C min minimum cost of locomotion - t end endurance time  相似文献   

18.
The half-saturation constant ( K 8) for growth and the maximum growth rate (μmax) were determined for 2 clones of Thalassiosira pseudonana (=Cyclotella nana) under conditions in which external silicon concentrations controlled growth. The estuarine clone (3H) had a higher half-saturation constant and maximum growth rate ( K 8= 0.98 μM Si; μmax= 3.6 divisions/day) than the clone from the Sargasso Sea ( K 8= 0.19 μM Si; μmax= 2.1 divisions/day). The K 8 values for each clone are such that the silicate levels found at certain times in both the Sargasso Sea and the coastal regions are rate limiting to growth, hence can be of significance to plant production and to species succession. The yield data are consistent with the concept that growth rate and cellular silicon content vary together in silicon-limited cultures.  相似文献   

19.
Abstract The germination of Sorghum bicolor seeds of 9 genotypes was tested at temperatures between 8°C and 48°C on a thermal gradient plate. Samples were tested from three regions of the panicle expected to differ in temperature during grain filling. Seeds of a tenth genotype, SPV 354, produced in controlled-environment glasshouses at different panicle temperatures, were tested similarly. In addition, the emergence of SPV 354 was measured from planting depths of 2 and 5 cm at mean soil temperatures of 15, 20 and 25°C. Four methods of calculating mean germination rate for the nine genotypes were compared. Germination characters like base, optimum and maximum temperature (Tb, To, Tm), thermal time (θ)and the germination rate at To(Rmax showed only small differences between methods. There was a range of genotypic variation in all characters: Tb 8.5–11.9°C; To, 33.2–37.5°C; Tm, 46.8–49.2°C; θ, 23.4–38.0°Cd; Rmax, 0.69–1.14-d-1. In contrast, mean germinability (G) was between 90% and 100% over the temperature range 13–40°C. Panicle temperature had no effect on any germination character in SPV 354. However, deeper burial increased θ for emergence and decreased G, irrespective of soil temperature except at 5 cm. Increasing panicle temperature, by reducing seed size, reduced G and increased θ by about 10% only at 15°C and 5 cm depth.  相似文献   

20.
The photosynthetic temperature response of the Antarctic vascular plants Colobanthus quitensis and Deschampsia antarctica was examined by measuring whole-canopy CO2 gas exchange and chlorophyll (Chl) a fluorescence of plants growing near Palmer Station along the Antarctic Peninsula. Both species had negligible midday net photosynthetic rates (Pn) on warm, usually sunny, days (canopy air temperature [Tc]> 20°C), but had relatively high Pn on cool days (Tc<10°C). Laboratory measurements of light and temperature responses of Pn showed that high temperature, not visible irradiance, was responsible for depressions in Pn on warm sunny days. The optimal leaf temperatures (Tl) for Pn in C. quitensis and D. antarctica were 14 and 10°C, respectively. Both species had substantial positive Pn at 0°C Tl, which were 28 (C. quitensis) and 32% (D. antarctica) of their maximal Pn, and we estimate that their low-temperature compensation points occurred at ?2°C Tl (C. quitensis) and ?3°C (D. antarctica). Because of the strong warming trend along the peninsula over recent decades and predictions that this will continue, we were particularly interested in the mechanisms responsible for their negligible rates of Pn on warm days and their unusually low high-temperature compensation points (i.e., 26°C in C. quitensis and 22°C in D. antarctica). Low Pn at supraoptimal temperature (25°C) appeared to be largely due to high rates of temperature-enhanced respiration. However, there was also evidence for direct impairment of the photosynthetic apparatus at supraoptimal temperature, based on Chl fluorescence and Pn/intercellular CO2 concentration (ci) response curve analyses. The breakpoint or critical temperature (Tcr) of minimal fluorescence (Fo) was ≈42°C in both species, which was well above the temperatures where reductions in Pn were evident, indicating that thylakoid membranes were structurally intact at supraoptimal temperatures for Pn. The optimal Tl for photochemical quenching (qp) and the quantum yield of photosystem II (PSII) electron transfer (φPSII) were 9 and 7°C in C. quitensis and D. antarctica, respectively. Supraoptimal temperatures resulted in lower qp and greater non-photochemical quenching (qNP), but had little effect on Fo, maximal fluorescence (Fm) or the ratio of variable to maximal fluorescence (Fv/Fm) in both species. In addition, carboxylation efficiencies or initial slopes of their Pn/ci response were lower at supraoptimal temperatures, suggesting reduced activity of ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco). Although continued warming along the peninsula will increase the frequency of supraoptimal temperatures, Tc at our field site averaged 4.3°C and was below the temperature optima for Pn in these species for the majority of diurnal periods (86%) during the growing season, suggesting that continued warming will usually improve their rates of Pn.  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