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 Using a physical model of the capacity for non-evaporative heat loss and measurements of metabolic heat production, I evaluated the evaporative requirements for thermoregulation in the deer mouse,Peromyscus maniculatus, and the white rabbit,Oryctolagus cuniculus. The physical limit to non-evaporative heat loss was calculated from the heat transfer properties of the two animals and expressed as a maximum thermal conductance (C max). Two physiologically-based thermal conductances were derived from evaporative water loss, respiratory gas exchange and core temperature measurements made between 8 and 34°C on the deer mouse, and taken from published data for the white rabbit. The thermal conductance for non-evaporative heat loss (C) was calculated from net heat production, whereasC m represented the thermal conductance required to dissipate metabolic heat production. Evaporation is required when metabolic heat production exceeds the capacity for non-evaporative heat loss (as shown byC m>C max). However, evaporation increased in both animals although additional capacity to lose heat remained (i.e.,C<C max). Evaporation increased withC above 30°C for the mouse and at each 5°C measurement interval from 15 to 30°C for the rabbit. Thus, evaporation was greater than that required for thermoregulation for both animals as determined from a physical model of heat loss because both evaporation andC increased together to regulate heat loss.Symbols: see list on title page of preceeding paper  相似文献   

2.
Summary The gas exchange characteristics of two C3 desert annuals with contrasting phenologies, Geraea canescens T. & G. (winter-active) and Dicoria canescens T. & G. (summer-active), both Asteraceae, were determined for plants grown under a moderate (25°/15° C, day/night temperature) and a high (40°/27° C) growth temperature regime. Both species had high photosynthetic capacities; maximum net photosynthetic rates were 38 and 48 mol CO2 m-2 s-1 for Geraea and Dicoria, respectively, and were not influenced by growth temperature regime. However, the temperature optima of net photosynthesis shifted from 26° C for Geraea and from 28° C for Dicoria when grown under the moderate temperature regime to 31° C for both species when grown under the high temperature regime. Although the shifts in temperature optima were smaller than those observed for many desert perennials, both species showed substantial increases in photosynthetic rates at high temperatures when grown at 40°/27° C. In general, the gas exchange characteristics of Geraea and Dicoria were very similar to each other and to those reported for other C3 desert annuals. Geraea and Dicoria experienced different seasonal patterns of change in several environmental variables. For Geraea, maximum daily air temperature (T a) increased from 24° to 41° C over its growing season while Dicoria experienced maximum T a at midseason (45° C). At points during their respective growing seasons when midday T a ranged between 35° and 40° C, leaf temperatures (T 1) of both species were below T a and, therefore, were closer to the photosynthetic temperature optima measured in the laboratory. Leaf conductances to water vapor (g 1) and water potentials () were high at these times, but later in their growing seasons Dicoria maintained high g 1 and while Geraea showed large decreases in these quantities. The ability of Dicoria to successfully growth through the hot, dry summers of the California deserts may be related to its ability to acquire the available water in locally mesic habitats.  相似文献   

3.
The optimal conditions for opening of stomata in detached epidermis of the Crassulacean Acid Metabolism (CAM) plant Kalanchoe daigremontiana were determined. Stomatal opening in CO2–free air was unaffected by light so subsequently all epidermal strips were incubated in the dark and in CO2–free air. Apertures were maximal after 3 h incubation and were significantly greater at 15° C than 25° C. Thus stomata in isolated epidermis of this species can respond directly to temperature. Stomatal opening was greatest when the incubating buffer contained 17.6 mol m–3 K+, but decreased linearly with increasing K+ concentrations between 17.6 and 300 mol m–3; the decrease in aperture was shown to be associated with increasing osmotic potentials of the solutions. Reasons for this behaviour, which differs from that of many C3 and C4 species, are discussed. Stomatal apertures declined linearly upon incubation of epidermis on buffer solutions containing between 10–11 and 10–5 mol m–3 abscisic acid (ABA). Hence stomata on isolated epidermis of K. daigremontiana respond to lower concentrations of ABA than those of any species reported previously.  相似文献   

4.
Homogeneity-time is defined and introduced as the criterion for mixing quality in bioreactors. The criterion could replace the mixing time, in the case, when more than one measuring point (sensors) is included in the measuring system. Results based on the homogeneity-time and the temperature pulse method, achieved in stirred tank reactors under aerated conditions as well as in a jet-mixed tank, are presented.List of Symbols C p,p kJ/kg K Heat capacity of the pulse medium - C p,s kJ/kg K Heat capacity of the reactor-medium - F m3/s Flow rate of the pulse-input - i Inhomogeneity - I N Inhomogeneity-number - M (t) °C Ideal response curve - m deNumber of combinations for certain number of sensors acc. to Table 1 - n Number of sensor - p kg/m3 Density of the pulse medium - kg/m3 Density of the tank medium - s 1 °C Mean absolute deviation of the sensor temperatures related on the ideal response curve s2 s Mean absolute deviation of the homogeneity-times related on the time achieved with 6 sensors - t s Time - t (i) s Homogeneity-time - t ps s Starting time of tracer injection - t PE s End time of tracer injection - T E °C Mean medium temperature at the end of experiment - T k °C Temperature at k-th sensor position - T p °C Pulse temperature - T s °C Mean medium temperature before the tracer injection - V s m3 Tank volume before pulse input  相似文献   

5.
Net photosynthetic rates and mesophyll conductances were measured under standardized conditions for leaves of two C3 and one C4 annual species grown at temperatures of 20 to 32°C. Plants were grown with varying day and night temperatures, and also at constant temperatures equal to all the day and night temperatures used. Plants were grown with 8, 12, and 16 hours of light per day. This design allowed determination of whether photosynthetic characteristics were best correlated with day, night, mean, or time-weighted mean temperatures, The results showed that for Glycine max (L.) Merr. (C3) night temperature was most important in determining photosynthetic characteristics, while in Helianthus annuus L. (C3) and Amaranthus hypochondriacus L. (C4) the time-weighted mean temperature was most important. The results for all species were consistent with the hypothesis that development of photosynthetic characteristics is related to a balance between the rate of leaf expansion and the rate of photosynthesis under the growth conditions.  相似文献   

6.
Thermoregulatory sweating [total body (m sw,b), chest (m sw,c) and thigh (m sw,t) sweating], body temperatures [oesophageal (T oes) and mean skin temperature (T sk)] and heart rate were investigated in five sleep-deprived subjects (kept awake for 27 h) while exercising on a cycle (45 min at approximately 50% maximal oxygen consumption) in moderate heat (T air andT wall at 35° C. Them sw,c andm sw,t were measured under local thermal clamp (T sk,1), set at 35.5° C. After sleep deprivation, neither the levels of body temperatures (T oes,T sk) nor the levels ofm sw, b,m sw, c orm sw, t differed from control at rest or during exercise steady state. During the transient phase of exercise (whenT sk andT sk,1 were unvarying), them sw, c andm sw, t changes were positively correlated with those ofT oes. The slopes of them sw, c versusT oes, orm sw, t versusT oes relationships remained unchanged between control and sleep-loss experiments. Thus the slopes of the local sweating versusT oes, relationships (m sw, c andm sw, t sweating data pooled which reached 1.05 (SEM 0.14) mg·cm–2·min–1°C–1 and 1.14 (SEM 0.18) mg·cm–2·min–1·°C–1 before and after sleep deprivation) respectively did not differ. However, in our experiment, sleep deprivation significantly increased theT oes threshold for the onset of bothm sw, c andm sw, t (+0.3° C,P<0.001). From our investigations it would seem that the delayed core temperature for sweating onset in sleep-deprived humans, while exercising moderately in the heat, is likely to have been due to alterations occurring at the central level.  相似文献   

7.
Summary The effect of doubling the saturated fatty acid content on the electrophysiology ofNeurospora crassa membranes was studied. Intracellular membrane input resistance (R m ) and potential (E m ) were measured for wild-type (w/t) andcel (Tween 40) organisms as a function of temperature. Over the 0 to 40°C temperature range studied, meanE m values of bothw/t andcel (Tw 40) organisms increased from –160 to –210 mV. This difference is greater than that expected from Nernst potential considerations, indicating an active component ofE m . This active component is insensitive to a doubling of the saturated fatty acid content.R m exhibits a temperature dependence and hysteresis. Averaged data indicate an increase inR m with decreased temperature. The slope of the temperature dependence varies among individual hyphae. Above 17.5°Ccel (Tw 40) hyphae averaged greater than 70% higher values ofR m thanw/t. Below 17.5°Cw/t R m data divided into low and high temperature dependence groups, whilecel data exhibited a low temperature dependence. The results are discussed in relation to gel-liquid crystal phase transitions, membrane fluidity, and the contribution of fatty acid structure to membrane electrical properties.  相似文献   

8.
Species in the Laxa and Grandia groups of the genus Panicum are adapted to low, wet areas of tropical and subtropical America. Panicum milioides is a species with C3 photosynthesis and low apparent photorespiration and has been classified as a C3/C4 intermediate. Other species in the Laxa group are C3 with normal photorespiration. Panicum prionitis is a C4 species in the Grandia group. Since P. milioides has some leaf characteristics intermediate to C3 and C4 species, its photosynthetic response to irradiance and temperature was compared to the closely related C3 species, P. laxum and P. boliviense and to P. prionitis. The response of apparent photosynthesis to irradiance and temperature was similar to that of P. laxum and P. boliviense, with saturation at a photosynthetic photo flux density of about 1 mmol m-2 s-1 at 30°C and temperature optimum near 30°C. In contrast, P. prionitis showed no light saturation up to 2 mmol m-2 s-1 and an optimum temperature near 40°C. P. milioides exhibited low CO2 loss into CO2-free air in the light and this loss was nearly insensitive to temperature. Loss of CO2 in the light in the C3 species, P. laxum and P. boliviense, was several-fold higher than in P. milioides and increased 2- to 5-fold with increases in temperature from 10 to 40°C. The level of dark respiration and its response to temperature were similar in all four Panicum species examined. It is concluded that the low apparent photorespiration in P. milioides does not influence its response of apparent photosynthesis to irradiance and temperature in comparison to closely related C3 Panicum species.Abbreviations AP apparent photosynthesis - I CO2 compensation point - gl leaf conductance; gm, mesophyll conductance - PPFD photosynthetic photon flux density - PR apparent photorespiration rate - RuBPC sibulose bisphosphate carboxylase  相似文献   

9.
Summary The effect of temperature on the ability of neuromuscular junctions and muscle fibers to contract, release neurotransmitter, and maintein postsynaptic membrane properties, was measured in vivo and in vitro in claw closer muscles in stone crabsMenippe mercenaria (Say) and blues crabsCallinectes sapidus (Rathbun).In vivo muscle stress (defined as force generated per unit of muscle cross-sectional area) was measured in crabs of both species collected from northern populations (annual temperature range 2–30°C) and southern populations (annual temperature range 15–30°C). Muscle stress was compared between (1) crabs of both species maintained in the laboratory at 30°C (laboratory warmed); (2) crabs given a brief acclimation period (4 weeks for blue crabs; 7 weeks for stone crabs) at 8°C in the laboratory (laboratory colled), and (3) stone crabs that had been naturally acclimated from summer (30°C) to winter (8°C) temperatures over a 6 month period in the field (naturally cooled). No differences were found in the abilities of the northern and southern populations of either species to generate muscle stress when tested at summer temperatures (30°C) common to both populations. Northern and southern blue crabs produced similar levels of muscle stress whether laboratory warmed (30°C) or laboratory cooled (8°C). Conversely, northern and southern stone crabs showed significantly reduced muscle stress in laboratory cooled crabs compared with laboratory warmed crabs. Stone crabs from both populations generated the same amount of muscle stress after having been naturally cooled to 8°C as they had generated the previous summer (30°C).In vitro neuromuscular properties (i.e. (1) muscle stress as a measure of contractile ability; (2) excitatory junction potential (EJP) amplitude as a measure of neurotransmitter release; (3) specific membrane resistance (Rm) as a measure of postsynaptic membrane properties) were compared at 8, 20, and 30°C between northern cold acclimated (naturally cooled stone crabs and laboratory cooled blue crabs) and non-cold acclimated (laboratory cooled stone crabs. Muscle fibers in claws of stone crabs and blue crabs showing cold acclimation had higher Rm at 8°C than non-cold acclimated crabs. This higher Rm resulted in a broadening of EJP's which enhanced EJP summation, muscle fiber depolarization, and muscle stress.Abbreviations EJP excitatory junction potential - E r resting membrane potential - F e lacilitation - R m specific membrane resistance  相似文献   

10.
Summary A special temperature control system has been developed and applied to continuous measuring of the heat evolved during a fermentation process. In this system, the fermentation broth was overcooled by a given constant cooling water flow. The excess heat removed from the fermentor was then made up by an immersion electrical heater. The action of the temperature controller was precisely monitored as it varied in response to the amount of heat produced by the microbial activities.The technique was used for determining the heat evolution byEscherichia coli grown on glucose. The ratio between quantities of total heat release and total oxygen consumption has been determined to be 0.556 MJ/mol O2.The newly developed technique can be employed as an online sensor to monitor the microbial activities of either aerobic or anaerobic fermentation systems.Symbols Cc Heat capacity of cooling water (MJ/kg · °C) - Cp Heat capacity (MJ/kg · °C) - I Current of immersion heater (A) - K Constant in Equation (2) (h) - K Constant in Equation (13) (m3 · h · °C/MJ) - Qc Flow rate of cooling water (m3/h) - Heat of agitation (MJ/m3 · h) - Heat dissipated by the bubbling gas (MJ/m3 · h) - Heat removal by the action of controller (MJ/m3 · h) - Heat of fermentation (MJ/m3 · h) - Heat loss to the surroundings (MJ/m3 · h) - Qpass Constant average power dissipated by the immersion heater (MJ/m3 · h) - Fluctuating power dissipated by the immersion heater (MJ/m3 · h) - Power dissipated by the immersion heater (MJ/m3 · h) - T Temperature of fermentation broth (°C) - Constant average temperature of fermentation broth (°C) - Fluctuating temperature of fermentation broth (°C) - Ta Temperature of the ambient air (°C) - Tc Inlet temperature of cooling water (°C) - U1A1 Specific heat transfer coefficient for determination of heat loss to the surroundings (MJ/m3 · h · °C) - U2A2 Specific heat transfer coefficient for cooling surfaces (MJ/m3 · h · °C) - U3A3 Constant in Equation (16) (MJ/m3 · h · °C) - V Voltage of immersion heater (V) - VL Liquid volume (m3) - OUR Oxygen uptake rate (mol O2/m3 · h) Greek Letters Hfo The ratio between the total heat release and the total oxygen uptake (MJ/mol O2) - c Density of cooling water (kg/m3) - Time constant defined in Equation (6) (h) - iMiCpi Heat capacity of system components (fermentation broth + fermentor jar + stainless steel) (MJ/m3 · °C)  相似文献   

11.
Summary The effect of temperature on the response properties of primary auditory fibres in caiman was studied. The head temperature was varied over the range of 10–35 ° C while the body was kept at a standard temperature of 27 °C (Ts). The temperature effects observed on auditory afferents were fully reversible. Below 11 °C the neural firing ceased.The mean spontaneous firing rate increased nearly linearly with temperature. The slopes in different fibres ranged from 0.2–3.5 imp s–1 °C–1. A bimodal distribution of mean spontaneous firing rate was found (<20 imp s–1 and >20 imp s–1 at Ts) at all temperatures.The frequency-intensity response area of the primary fibres shifted uniformly with temperature. The characteristic frequency (CF) increased nearly linearly with temperature. The slopes in different fibres ranged from 3–90 Hz °C–1. Expressed in octaves the CF-change varied in each fibre from about O.14oct °C–1 at 15 °C to about 0.06 oct °C–1 at 30 °C, irrespective of the fibre's CF at Ts. Thresholds were lowest near Ts. Below Ts the thresholds decreased on average by 2dB°C–1, above Ts the thresholds rose rapidly with temperature. The sharpness of tuning (Q10db) showed no major change in the temperature range tested.Comparison of these findings with those from other lower vertebrates and from mammals shows that only mammalian auditory afferents do not shift their CF with temperature, suggesting that a fundamental difference in mammalian and submammalian tuning mechanisms exists. This does not necessarily imply that there is a single unifying tuning mechanism for all mammals and another one for non-mammals.Abbreviations BF best frequency: frequency of maximal response at an intensity 10 dB above the CF-threshold - CF characteristic frequency - FTC frequency threshold curve, tuning curve - T s standard temperature of 27 °C  相似文献   

12.
Human red cell membrane bindings of arachidonate and palmitate at pH 7.3 are investigated at temperatures between 0 and 38°C by equilibrating ghosts with the long-chain fatty acids bound to bovine serum albumin in molar ratios (v) within the physiological range (<1.7). Linearized relations of ghost uptakes and fatty acid monomer concentrations in buffer provide estimates of the binding capacities and corresponding equilibrium dissociation constants (K dm ). The temperature-independent arachidonate binding capacity, 5.5 ± 0.5 nmol g–1 packed ghosts, is approximately fivefold smaller than that of palmitate, 26.6 ± 2.0 nmol g–1. While K dm of arachidonate binding 5.1 ± 0.5 nm is temperature independent, K dm of palmitate increases with temperature from 3.7 nm at 0°C to 12.7 nm at 38°C.The large difference in binding capacities suggests the presence of at least two different fatty acid binding domains in human red cell membranes.  相似文献   

13.
We present a model for a conditional bursting neuron consisting of five conductances: Hodgkin-Huxley type time- and voltage-dependent Na+ and K+ conductances, a calcium activated voltage-dependent K+ conductance, a calcium-inhibited time- and voltage-dependent Ca++ conductance, and a leakage Cl( conductance. With an initial set of parameters (versionS), the model shows a hyperpolarized steady-state membrane potential at which the neuron is silent. Increasingg Na and decreasingg Cl, whereg i , is the maximal conductance for speciesi, produces bursts of action potentials (BursterN). Alternatively, an increase ing Ca produces a different bursting state (BursterC). The two bursting states differ in the periods and amplitudes of their bursting pacemaker potentials. They show different steady-stateI–V curves under simulated voltage-clamp conditions; in simulations that mimic a steady-stateI–V curve taken under experimental conditions only BursterN shows a negative slope resistance region. ModelC continues to burst in the presence of TTX, while bursting in ModelN is suppressed in TTX. Hybrid models show a smooth transition between the two states.  相似文献   

14.
The intrinsic rate of natural increase(rm) is useful to estimate the populationgrowth potential of insects and mites, whichmay help predict the outcome of pest-naturalenemy interactions. This study was conductedto determine how 12 constant temperatureregimes between 10–38 °C (± 0.5 °C) may differentially affect therm of the McDaniel spider mite, Tetranychus mcdanieli McGregor (Acarina: Tetranychidea), a common pest of raspberry, andits coccinellid predator, Stethoruspunctillum Weise (Coleoptera: Coccinellidae). Tetranychus mcdanieli survived tomaturity in the 14–36 °C range, comparedto the 14–32 °C range for S.punctillum. Survival above 24 °Cremained high for the spider mite, butdecreased markedly for the coccinellid. Tetranychus mcdanieli's range forreproduction was similar to its survival range,but S. punctillum failed to reproduce at14 °C and reproduced only poorlyat 16 °C. Offspring production peakedat 24 °C for both T. mcdanieli(average 152 eggs per female), and S.punctillum (280 eggs per female). At alltemperatures suitable for reproduction, femalelongevity of the coccinellid was greater thanthe spider mite, which was characterized byearlier/faster reproduction than thecoccinellid. As temperature increased, therm followed a typical asymmetricaldome-shape pattern, with maximum values of0.196 d–1 and 0.385 d–1at 30 °C and 34 °C, for S.punctillum and T. mcdanieli,respectively. For each species, therm-temperature relation was successfullymodelled using a curvilinear regressionequation previously shown to predictdevelopment rate. In both species, thedevelopment rate response to temperature has amajor influence on the temperature-rmrelationship. In the 16–32 °C rangesuitable for population growth of both species,the rm of T. mcdanieliwas 1.9 (30 °C) to 8 (16 °C) times greaterthan S. punctillum. These growthpotential ratios are consistently in favor ofthe prey, suggesting a limitation of thecoccinellid with respect to its capacity totrack T. mcdanieli populations. However,under short season conditions, the inferiorreproductive dynamics of S. punctillum'svs. spider mite prey should not have aprevailing influence in determining impact, andmay be compensated by high voracity incombination with a strong aggregativeresponse.  相似文献   

15.
Summary The influence of light intensity and temperature on the diurnal course and magnitude of CO2 gas exchange and on acid metabolism was studied in the laboratory with rooted rosettes of Sempervivum montanum collected at 2,200 m above sea level in the Central Alps. Under a temperature regime having a cool dark period and warm light period, S. montanum exhibited the time course of CO2 gas exchange typical of a CAM plant; the response was very distinct even when the plants were well-watered. At day temperatures of less than 10° C and at night temperatures greater than 35° C, S. montanum behaved like a C3 plant. Characteristic for S. montanum are a broad temperature optimum and a wide range of temperatures in which CO2 uptake in light is possible (-2° to 45° C). Dark fixation of CO2 is evident between-2° and 35° C, an apparent uptake of external CO2, on the other hand, only as high as 20° C. Light saturation of CO2 uptake is reached at 60–80 W m-2 while the rate of deacidification is nearly maximal at 40 W m-2. These results show that, due to their specific metabolism, CAM plants can be favored not only in xeric habitats, but also in heat stressed mountain habitats where the daily variation in temperature may be extreme.Dedicated with appreciation to Dr. K.F. Springer  相似文献   

16.
J. Silvola  S. Heikkinen 《Oecologia》1979,37(3):273-283
Summary The CO2 exchange of the Empetrum nigrum-Sphagnum fuscum community of a raised bog was studied in the laboratory at different temperature (from 5 to 30° C) and irradiance (up to 128 W m-2) combinations during one growing season. The total CO2 exchange was divided into three components, namely those due to Empetrum nigrum, Sphagnum fuscum, and peat, respectively. At the optimum temperature (10 to 15° C) the maximum net CO2 exchange of Empetrum nigrum was c. 200 and that of Sphagnum fuscum c. 250 mg CO2 m-2h-1. The total respiration in peat increased exponentially from 50 to 350 mg CO2 m-2h-1 with increasing temperature from 5 to 30° C. About 40% of the CO2 fixed by the community in optimal temperature and irradiation conditions was released immediately.  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 effect of leaf temperature, O2 and calculated O2/CO2 solubility ratio in the leaf on the quantum yield of photosynthesis was studied for the C4 species, Zea mays L., and the C3 species, Triticum aestivum L. Over a range of leaf temperatures of 16 to 35° C, the quantum yield of Z. mays was relatively constant and was similar under 1.5 and 21% O2, being ca. 0.059 mol CO2 mol-1 quanta absorbed. Under 1.5% O2 and atmospheric levels of CO2, the quantum yield of T. aestivum was relatively constant (0.083 mol CO2 mol-1 quanta absorbed) at leaf temperatures from 15 to 35° C. Atmospheric levels of O2 (21%) reduced the quantum yield of photosynthesis in T. aestivum and as leaf temperature increased, the quantum yield decreased from 0.062 at 15°C to 0.046 mol CO2 mol-1 quanta absorbed at 35°C. Increasing temperature decreases the solubility of CO2 relatively more than the solubility of O2, resulting in an increased solubility ratio of O2/CO2. Experimentally manipulating the atmospheric levels of O2 or CO2 to maintain a near-constant solubility ratio of O2/CO2 at varying leaf temperatures largely prevented the temperature-dependent decrease in quantum yield in t. aestivum. Thus, the decrease in quantum yield with increasing leaf temperature in C3 species may be largely caused by a temperaturedependent change in the solubility ratio of O2/CO2.J and II=Ku and Edwards, 1977a, b  相似文献   

19.
Divalent cation (Mn2+, Ca2+) entry into rat parotid acinar cells is stimulated by the release of Ca2+ from the internal agonist-sensitive Ca2+ pool via a mechanism which is not yet defined. This study examines the effect of temperature on Mn2+ influx into internal Ca2+ pool-depleted acini (depl-acini, as a result of carbachol stimulation of acini in a Ca2+-free medium for 10 min) and passive 45Ca2+ influx in basolateral membrane vesicles (BLMV). Mn2+ entry into deplacini was decreased when the incubation temperature was lowered from 37 to 4°C. At 4°C, Mn2+ entry appeared to be inactivated since it was not increased by raising extracellular [Mn2+] from 50 m up to 1 mm. The Arrhenius plot of depletion-activated Mn2+ entry between 37 and 8°C was nonlinear, with a change in the slope at about 21°C. The activation energy (Ea) increased from 10 kcal/mol (Q10=1.7) at 21–37°C to 25 kcal/mol (Q10=3.0) at 21-8°C. Under the same conditions, Mn2+ entry into basal (unstimulated) cells and ionomycin (5 m) permeabilized depl-acini exhibit a linear decrease, with E a of 7.8 kcal/mol (Q10=1.5) and 6.2 kcal/mol (Q10 < 1.5), respectively. These data suggest that depletion-activated Mn2+ entry into parotid acini is regulated by a mechanism which is strongly temperature dependent and distinct from Mn2+ entry into unstimulated acini.As in intact acini, Ca2+ influx into BLMV was decreased (by 40%) when the temperature of the reaction medium was lowered from 37 to 4°C. Kinetic analysis of the initial rates of Ca2+ influx in BLMV at 37°C demonstrated the presence of two Ca2+ influx components: a saturable component, with K Ca =279 ± 43 m, Vmax = 3.38 ± 0.4 nmol Ca2+/mg protein/min, and an apparently unsaturable component. At 4°C, there was no significant change in the affinity of the saturable component, but Vmax decreased by 61% to 1.3 ± 0.4 nmol Ca2+/mg protein/min. There was no detectable change in the unsaturable component. When BLMV were treated with DCCD (5 mm) or trypsin (1100, enzyme to membrane) for 30 min at 37°C there was a 40% decrease in Ca2+ influx. When BLMV were treated with DCCD or trypsin at 4°C and subsequently assayed for Ca2+ uptake at 37°C there was no significant loss of Ca2+ influx. These data suggest that the temperature sensitive high affinity Ca2+ flux component in BLMV is mediated by a protein which undergoes a modification at low temperatures, resulting in decreased Ca2+ transport.We thank Dr. Bruce Baum, Dr. Yukiharu Hiramatsu, Dr. Ofer Eidelman, and our other colleagues for their support during this work.  相似文献   

20.
Physical work capacity (PWC180) was assessed with different levels of hypohydration in 25 heat-acclimatized male volunteers in hot dry (45°C DB, 30% RH) and hot humid (39°C DB, 60% RH) conditions equated to a heat stress level of 34°C on the WBGT scale. Heat acclimatization was carried out by exposing the subjects for 8 consecutive days in a climatic chamber with moderate work for two 50 min work cycles and 10 min intervening rest pauses. Acclimatization resulted in significant decreases in heart rate (27 bpm), oral temperature (0.8°C), mean skin temperature (1.2°C) and a significant increase in sweating rate (120 g h–1 m–2). Day-to-day variations in body hypohydration levels during heat acclimatization were not significantly different, although water intake was found to increase significantly from day 3 onwards when the subjects were in ad lib water intake state. The heat acclimatized subjects were then hypohydrated to varying degrees, viz. 1%, 2% and 3% body weight deficit, with moderate work in heat in the climatic chamber and after successful recovery from the effects of thermal stress and exercise; their physical work capacity was assessed individually. Physical work capacity was found to decrease significantly with hypohydration as compared to controls. The decrease was of the order of 9%, 11% and 22% in the hot dry condition and 6%, 8% and 20% in the hot humid condition with hypohydration levels of 1%, 2% and 3% respectively. The decrease was more pronounced during 3% hypohydration level under both heat stress conditions. This decrease was in spite of significant increases in maximal ventilation. However, the PWC180 under the two heat stress conditions, when compared, did not reveal any significant difference. It was concluded that the heat stress vehicle did not adversely affect the physical work capacity. On the other hand, the decreases in physical work capacity were found to be closely related to the primary hypohydration level in heat-acclimatized tropical subjects.Abbreviations WBGT wet bulb globe temperature - bam beats per minute - YSI Yellow Springs Instrument - EKG electrocardiogram  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