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 46 毫秒
1.
Fourteen adult patients undergoing open heart surgery under induced hypothermia had median nerve, short-latency somatosensory evoked potentials (SSEPs) recorded during cooling (from 36°C to 19°C) and subsequent rewarming. Similar data on another group of patients who had brain-stem auditory evoked potentials (BAEPs) were also analyzed. Hypothermia produced increased latencies of the major SSEP and BAEP components and the latencies returned to normal with subsequent warming. The temperature-latency relationship during the cooling phase was significantly different from that during the warming phase. For SSEP components the temperature-latency relationship was linear during cooling and curvilinear during warming, whereas for BAEP it was curvilinear both during cooling and warming. Furthermore, the regression curves were different during the two phases of temperature manipulation, particularly for temperatures below 30°C both for SSEP and BAEP components. At the onset of warming there was an initial exaggerated warming response on the evoked potential (EP) latencies and amplitude of the EP components. The temperature-latency regression curves were uniformly less steep during the warming phase compared to those during cooling. These findings suggest the existence of hysteresis in the relationship between temperature and EP latencies. The latencies at a given temperature below 30°C depend on whether that temperature is reached during cooling or during warming.  相似文献   

2.
Due to their ecological, physiological, and molecular adaptations to low and varying temperatures, as well as varying seasonal irradiances, polar non-marine eukaryotic microalgae could be suitable for low-temperature biotechnology. Adaptations include the synthesis of compounds from different metabolic pathways that protect them against stress. Production of biological compounds and various biotechnological applications, for instance, water treatment technology, are of interest to humans. To select prospective strains for future low-temperature biotechnology in polar regions, temperature and irradiance of growth requirements (Q10 and Ea of 10 polar soil unicellular strains) were evaluated. In terms of temperature, three groups of strains were recognized: (i) cold-preferring where temperature optima ranged between 10.1 and 18.4°C, growth rate 0.252 and 0.344 · d−1, (ii) cold- and warm-tolerating with optima above 10°C and growth rate 0.162–0.341 · d−1, and (iii) warm-preferring temperatures above 20°C and growth rate 0.249–0.357 · d−1. Their light requirements were low. Mean values Q10 for specific growth rate ranged from 0.7 to 3.1. The lowest Ea values were observed on cold-preferring and the highest in the warm-preferring strains. One strain from each temperature group was selected for PN and RD measurements. The PN:RD ratio of the warm-preferring strains was less affected by temperature similarly as Q10 and Ea. For future biotechnological applications, the strains with broad temperature tolerance (i.e., the group of cold- and warm-tolerating and warm-preferring strains) will be most useful.  相似文献   

3.
Participants of the Chernobyl clean-up (n = 145) teams exposed to radiation doses from 0.05 to 3.5 Gy who had for the first time complained of pathologic somatosensory sensations (ostealgic syndrome), 20 healthy subjects, and 50 veterans of the war in Afghanistan with posttraumatic stress disorder (PTSD) were examined by a neuropsychiatrist and presented with the MMPI test. Somatosensory evoked potentials (SSEPs) were recorded. Paresthesia and cenesthopathy were characteristic of the participants of the Chernobyl clean-up. Sensation disorders of the cerebral type, kinesthetic illusions, cenesthopathic hypochondriac disorders, and paroxysmal psychosensory states predominated in this group of subjects. They differed significantly from the veterans with PTSD in markedly increased scores on MMPI scales (hypochondriasis, schizophrenia, pure hypochondriasis, pure schizophrenia, emotional exclusion, and perception oddity), which closely correlated with clinical somatosensory symptoms. In clean-up workers, somatosensory disorders were significantly associated with hypochondriac and schizophrenic-like symptoms. The latencies (LPs) of main SSEP components—N 20, P 25, N 140, P 300, and N 400—were increased and their amplitudes decreased in subjects exposed to radiation. Their SSEPs had significant topographical deviations in the left temporoparietal area: the contralateral LPs were increased, whereas the contralateral amplitudes of the thalamocortical N 20 component and the cortical P 25 component were decreased as compared to normal values. Somatosensory disorders and hypochondriac and schizophrenic symptoms were significantly correlated with changes in the SSEPs. The decrease in the N 20 amplitude and increase in the P 25 latency in the left temporoparietal area were dose-dependent. The results suggest cerebral rather than peripheral origin of ostealgic syndrome and other somatosensory disorders in the participants of the Chernobyl clean-up. These disorders are associated with radiation-induced dysfunction of the corticolimbic structures of the left—dominant—hemisphere. It is suggested that somatosensory disorders in patients exposed to low doses of radiation can be considered as manifestations of chronic fatigue syndrome /fibromyalgia, whereas schizoform organic brain lesions manifest themselves after exposure to a radiation dose of 0.3–0.5 Gy.  相似文献   

4.
To assess the role of different mechanisms in increasing the amplitude of the early components of cortical somatosensory evoked potentials (SSEPs) in lesions of central structures of the skin-motor analyzer in humans, SSEPs of the hand cortical projection zones (the points C3 and C4) and the spinal dorsal column nuclei (DCN) were recorded in parallel in response to trancutaneous electrostimulation of the median nerve in the carpal region in two groups of subjects. The control group included 26 healthy volunteers aged 39–62 years; the other group included 12 patients aged 45–63 years with hemiplegia and sensory disorders due to a stroke experienced 8–24 months before the electrophysiological studies. A significant (from P < 0.05 to P < 0.01) increase in the amplitude of the early SSEP components of the intact hemisphere and several early SSEP components of the affected hemisphere (with a decrease in the amplitude of the other components) and no changes in DCN SSEPs were observed in the patients compared to the control group, which was interpreted as a manifestation of local mechanisms causing an isolated increase in cortical excitability without changes in the reactivity of DCN.  相似文献   

5.
Four temperature treatments were studied in the climate controlled growth chambers of the Georgia Envirotron: 25/20, 30/25, 35/30, and 40/35 °C during 14/10 h light/dark cycle. For the first growth stage (V3-5), the highest net photosynthetic rate (P N) of sweet corn was found for the lowest temperature of 28–34 μmol m−2 s−1 while the P N for the highest temperature treatment was 50–60 % lower. We detected a gradual decline of about 1 P N unit per 1 °C increase in temperature. Maximum transpiration rate (E) fluctuated between 0.36 and 0.54 mm h−1 (≈5.0–6.5 mm d−1) for the high temperature treatment and the minimum E fluctuated between 0.25 and 0.36 mm h−1 (≈3.5–5.0 mm d−1) for the low temperature treatment. Cumulative CO2 fixation of the 40/35 °C treatment was 33.7 g m−2 d−1 and it increased by about 50 % as temperature declined. The corresponding water use efficiency (WUE) decreased from 14 to 5 g(CO2) kg−1(H2O) for the lowest and highest temperature treatments, respectively. Three main factors affected WUE, P N, and E of Zea: the high temperature which reduced P N, vapor pressure deficit (VPD) that was directly related to E but did not affect P N, and quasi stem conductance (QC) that was directly related to P N but did not affect E. As a result, WUE of the 25/20 °C temperature treatment was almost three times larger than that of 40/35 °C temperature treatment.  相似文献   

6.
A technique based on homogenisation of rapidly frozen tissue was used to investigate the regulation of intracellular pH (pHi) in freshwater and marine fish from diverse environmental temperatures. The following species were held at ambient temperatures of ca. 1°C (Notothenia coriiceps; Antarctica), 5°C (Pleuronectes platessa, Myoxocephalus scorpius; North Sea), and 26°C (Oreochromis niloticus; African lakes). The effects of seasonal acclimatisation to 4, 11 and 18°C were also examined in rainbow trout in the winter, autumn and summer, respectively. Extracellular (whole blood) pH (pHe) did not follow the constant relative alkalinity relationship, where pH+=pOH for any particular temperature, over a range of 1–26°C (overall δpHeT=0.009±0.002 U °C−1; P<0.001), apparently being regulated by ionic fluxes and ventilation. Intracellular pH (pHi) was also regulated independently of pN(=0.5 pK water) in all species of fish examined. The inverse relationship between pHi and environmental temperature gave an overall δpHiT of −0.010±0.001 U °C−1 (for both white and red muscle) and −0.004±0.003 U °C−1 (cardiac muscle). However, between 1 and 11°C δpHiT was much higher (P<0.001), −0.022±0.003 U °C−1 (white muscle) and −0.022±0.004 U °C−1 (red muscle). The possible adaptive roles for these different acid–base responses to environmental temperature variation among tissues and species, and the potential difficulties of estimating pHi, are discussed.  相似文献   

7.
S K Orme  G A Kelly 《Life sciences》1977,20(4):597-608
Although hypothermic whole organ perfusion is widely used in attempts to preserve organs for transplantation and to preserve the myocardium during cardiac surgery, little is known about substrate metabolism during hypothermia. A knowledge of metabolism utilization during hypothermic whole organ perfusion might allow optimal substrate choice for preservation of energy stores and functional capacity. Separate groups of hearts from fed rats were perfused 30 minutes with Krebs Henseleit bicarbonate buffer containing 5mM glucose-U-14C, at 37°, 25°, 20°, 15° and 10°C. From 37° to 15°C, heart rate decreased 90% and coronary flow decreased 25%. Glucose uptake decreased 5 fold from 37° to 10°C while 14CO2 and lactate production decreased 50 fold and 28 fold, respectively. Myocardial glycogen was stable until 10°C at which point increased glycogenolysis occured. The incorporation of 14C in glycogen was stable at 37°, 30° and 25° but decreased progressively with lower temperatures. The percent recovery of glucose as 14CO2, lactate and 14C in glycogen decreased from 73% at 37° at 10°C. Our studies indicate that metabolism of glucose is greatly reduced but significant above 15°C.  相似文献   

8.
The aim of this study was to evaluate the effects of low air temperature during nocturnal (TN) and diurnal (TD) periods as well as the substrate temperature (TS) on photosynthesis of ‘Valencia’ orange tree grafted on Rangpur lime rootstock. The experiment was carried out in a growth chamber with seven-month-old plants. The plants were exposed to the following temperature regimes: low substrate temperature (LTS, with: TD = 28°C, TN = 20°C, TS = 10°C); low air temperature during night (LTN, with: TD = 28°C, TN = 10°C, TS = 26°C); low temperature during nighttime and also low substrate temperature (LTSN, with: TD = 28°C, TN = 10°C, TS = 10°C); low air temperature during both diurnal and nocturnal periods (LTND, with: TD = 17°C, TN = 10°C, TS = 26°C); and finally to low air temperature (night and day) and low substrate temperature (LTSND, with: TD = 17°C, TN = 10°C, TS = 10°C). As reference (control), plants were subjected to TD = 28°C, TN = 20°C, and TS = 26°C. Measurements of leaf gas exchange, photochemical activity and carbohydrate concentrations were performed after six days of exposure to each thermal treatment. Compared to the control, all thermal regimes caused reductions in photosynthesis due to diffusive and metabolic limitations. The photoinhibition was transient in plants exposed to night and substrate low temperatures, whereas it was severe and chronic in plants subjected to chilling during the diurnal period. However, the lowest photosynthesis was observed in plants with low substrate temperature of 10°C (in LTS, LTSND and LTSN treatments), regardless of air temperature. The occurrence of cold night and/or its combination with low substrate temperature caused accumulation of starch in leaves. When considering carbohydrate concentrations in stems and roots, it was not possible to establish a clear response pattern to chilling. In conclusion, the low substrate temperature causes a greater reduction of CO2 assimilation in citrus plants as compared to the occurrence of low air temperature, being such response a consequence of diffusive and biochemical limitations.  相似文献   

9.
For the high production of phenylalanine by Escherichia coli, we cloned the pheAFR and aroFFR genes (FR = feedback resistant), which encoded chorismate mutase P-prephenate dehydratase and 3-deoxy-d-arabinoheptulosonate-7-phosphate synthase that are feedback inhibition-free as to the endproducts, into a temperature-controllable expression vector composed of the PR and PL promoter and a temperature sensitive repressor, cI857, of bacteriophage lambda. The plasmid obtained was designated as pSY130-14, and the temperature dependency of expression of the cloned genes and of phenylalanine production was investigated at different temperatures between 30 and 42°C using the strain AT2471 harbouring the plasmid. Above 35°C, the pheAFR gene and aroFFR gene expressions, and activities of both enzymes continued to increase up to 42°C. The cell concentration remained constant up to 38.5°C, but started to decrease sharply above 40°C, while the cell concentration of the host strain, AT2471, remained constant at all temperatures tested. The concentration of phenylalanine also depended on the temperature, and the highest production of phenylalanine, 18.6 g l−1, was obtained from glucose at 38.5°C in a 2.5 1 reactor.  相似文献   

10.
Grazing ruminants urinate and deposit N onto pastoral soils at rates up to 1,000 kg ha?1, with most of this deposited N present as urea. In urine patches, nitrous oxide (N2O) emissions can increase markedly. Soil derived CO2 fluxes can also increase due to priming effects.While N2O fluxes are affected by temperature, no studies have examined the interaction of pasture plants, urine and temperature on N2O fluxes and the associated CO2 fluxes. We postulated the response of N2O emissions to bovine urine application would be affected by plants and temperature. Dairy cattle urine was collected, labelled with 15N, and applied at 590 kg N ha?1 to a sub-tropical soil,with and without pasture plants at 11°, 19°, and 23°C. Over the experimental period (28 days), 0.2% (11°C with plants) to 2.2% (23°C with plants) of the applied N was emitted as N2O. At 11°C, plants had no effect on cumulative N2O-N fluxes, whereas at 23°C, the presence of plants significantly increased the flux, suggesting plant-derived C supply affected the N2O producing microbes. In contrast, a significant urine application effect on the cumulative CO2 flux was not affected by varying temperature from 11?C23°C or by growing plants in the soil. This study has shown that plants and their responses to temperature affect N2O emissions from ruminant urine deposition. The results have significant implications for forecasting and understanding the effect of elevated soil temperatures on N2O emissions and CO2 fluxes from grazed pasture systems.  相似文献   

11.
Temperature variation affects the growth, maturation and distribution of fish species due to increasing constraints on physiological functions therefore, the aim of the present study is to evaluate effect of temperature on thermal tolerance and standard metabolic rate (SMR) of gilthead seabream (Sparus aurata). For this purpose, tolerable temperature ranges of juvenile gilthead seabream acclimated at 15, 20, 25, and 30 °C for 30 days were estimated using dynamic and static thermal methodologies. The SMRs of the fish were also determined based on oxygen consumption rate (OCR). The dynamic and static thermal tolerance zones of gilthead seabream were calculated as 737 °C2 and 500 °C2, respectively, with a resistance zone area of 155.5 °C2. The SMR of the fish at the above acclimation temperatures (AT) was determined as 138, 257, 510, and 797 mg O2 h−1 kg−1, respectively and were significantly different (P < 0.01, n = 10). The temperature quotient (Q10) in relation to the SMR of the fish was calculated as 3.45, 3.91, and 2.44 for acclimation temperature ranges of 15–20, 20–25, and 25–30 °C, respectively. The fact that the SMR increased with rising temperatures and then decreased gradually after 25 °C indicates that the temperature preference of juvenile gilthead seabream lies between 25 and 30 °C. This study shows that gilthead seabream tolerates a relatively narrow temperature range, and consequently, a low capacity for acclimatisation to survive in aquatic systems characterised by temperature variations.  相似文献   

12.
《BBA》1986,851(2):267-275
The glow curve of chloroplasts excited by continuous light of high intensity (500 W · m−2) at pH 7.5 during cooling from +2 to −80°C consisted of seven bands appearing at about −30°C (TL−30), −15°C (TL−10), +10°C (TL+10), +30°C (TL+30), +50°C (TL+50), +65°C (TL+65) and +85°C (TL+80), in which TL stands for thermoluminescence. In the pH range from 5.5 to 9.0 the peak positions of the TL−30, TL−10, TL+50, TL+65 and TL+80 bands were independent of pH. On the other hand the peak positions of the TL+10 and TL+30 bands were gradually shifted from +25 to −5°C and from +20 to +40°C, respectively, as the pH was decreased from 9.0 to 5.5. The same pH-induced shift (from +25 to −5°C) was observed for the TL+10 band when electron transport was inhibited by DCMU. In dinoseb-treated chloroplasts the peak position of the main thermoluminescence band also exhibited pH dependency, and shifted from +20 to −20°C upon lowering the pH from 9.0 to 5.5. After the water-splitting system had been inactivated by Tris or NH2OH treatment no pH-induced shifts were observed in the peak positions of the thermoluminescence bands of DCMU and dinoseb-treated chloroplasts. The results suggest that the effect of pH on the thermoluminescence of untreated and inhibitor-treated chloroplasts is associated with protonation/deprotonation reactions occurring at the donor and acceptor sides of Photosystem II during the S1 → S2 transition of the water-splitting system.  相似文献   

13.
The speckled peacock bass Cichla temensis is a popular sport and food fish that generates substantial angling tourism and utilitarian harvest within its range. Its popularity and value make this species important for management and a potential aquaculture candidate for both fisheries enhancement and food fish production. However, little is known of optimal physiochemical conditions in natural habitats, which also are important for the development of hatchery protocols for handling, spawning and grow-out. Speckled peacock bass have been documented to have high sensitivity to extreme temperatures, but the metabolic underpinnings have not been evaluated. In this study, the effects of temperature (25, 30 and 35°C) on the standard metabolic rate (SMR) and lower dissolved oxygen tolerance (LDOT) of juvenile speckled peacock bass (mean ± standard error total length 153 ± 2 mm and wet weight 39.09 ± 1.37 g) were evaluated using intermittent respirometers after an acclimation period of 2 weeks. Speckled peacock bass had the highest SMR at 35°C (345.56 ± 19.89 mgO2 kg−1 h−1), followed by 30°C (208.16 ± 12.45 mgO2 kg−1 h−1) and 25°C (144.09 ± 10.43 mgO2 kg−1 h−1). Correspondingly, the Q10, or rate of increase in aerobic metabolic rate (MO2) relative to 10°C, for 30–35°C was also greater (2.76) than from 25 to 30°C (2.08). Similarly, speckled peacock bass were the most sensitive to hypoxia at the warmest temperature, with an LDOT at pO2 of 90 mmHg (4.13 mg l−1) at 35°C compared to pO2 values of 45 mmHg (2.22 mg l−1) and 30 mmHg (1.61 mg l−1) at 30 and 25°C, respectively. These results indicate that speckled peacock bass are sensitive to temperatures near 35°C, therefore we recommend managing and rearing this species at 25–30°C.  相似文献   

14.
The anation reaction of aquopentaamminecobalt(III) by acetate has been studied in the temperature range 60–80°C and acidity range 1.0 ≦ pH ≦ 5.5 for total acetate concentrations ≦ 0.5 M and at ionic strength 1.0 M. The anation by acetic acid follows second-order kinetics (k0), whereas the kinetic results for the anation by acetate (Q1, k1) provide evidence for the formation of an ion-pair with the complex ion. Typical experimental results at 70°C are k0 = 5.33 X 10−5 M−1 sec−1, Q1 = 5.87 M−1 and k1 = 1.46 X 10−4sec−1. The activation parameters for the different reaction paths are reported and the results discussed with reference to various other anation reactions of Co(III) complexes.  相似文献   

15.
Stands of carrot (Daucus carota L.) were grown in the field within polyethylene-covered tunnels at a range of soil temperatures (from a mean of 7·5°C to 10·9°C) at either 348 (SE = 4·7) or 551 (SE = 7·7) μmol mol−1 CO2. The effect of increased atmospheric CO2 concentration on root yield was greater than that on total biomass. At the last harvest (137d from sowing), total biomass was 16% (95% CI = 6%, 27%) greater at 551 than at 348 μmol mol−1 CO2, and 37% (95% CI = 30%, 44%) greater as a result of a 1°C rise in soil temperature. Enrichment with CO2 or a 1°C rise in soil temperature increased root yield by 31% (95% CI = 19%, 45%) and 34% (95% CI = 27%, 42%), respectively, at this harvest. No effect on total biomass or root yield of an interaction between temperature and atmospheric CO2 concentration at 137 DAS was detected. When compared at a given leaf number (seven leaves), CO2 enrichment increased total biomass by 25% and root yields by 80%, but no effect of differences in temperature on plant weights was found. Thus, increases in total biomass and root yield observed in the warmer crops were a result of the effects of temperature on the timing of crop growth and development. Partitioning to the storage roots during early root expansion was greater at 551 than at 348 μmol mol−1 CO2. The root to total weight ratio was unaffected by differences in temperature at 551 μmol mol−1CO2, but was reduced by cooler temperatures at 348 μmol mol−1 CO2. At a given thermal time from sowing, CO2 enrichment increased the leaf area per plant, particularly during early root growth, primarily as a result of an increase in the rate of leaf area expansion, and not an increase in leaf number.  相似文献   

16.
Net CO2 uptake rates (P N) were measured for the vine cacti Hylocereus undatus and Selenicereus megalanthus under relatively extreme climatic conditions in Israel. Withholding water decreased rates and the daily amount of CO2 uptake by about 10 % per day. Compared with more moderate climates within environmental chambers, the higher temperatures and lower relative humidity in the field led to a more rapid response to drought. The upper envelopes of scatter diagrams for P N versus temperature for these Crassulacean acid metabolism species, which indicate the maximal rates at a particular temperature, were determined for both night time CO2 uptake in Phase I (mediated by phosphoenolpyruvate carboxylase, PEPC) and early morning uptake in Phase II (mediated by ribulose-1,5-bisphosphate carboxylase/oxygenase, RuBPCO). As stem temperature increased above 13 °C, the maximal P N increased exponentially, reaching maxima near 27 °C of 12 and 8 μmol m−2 s−1 for Phases I and II, respectively, for H. undatus and 6 and 4 μmol m−2 s−1, respectively, for S. megalanthus. Based on the Arrhenius equation, the apparent activation energies of PEPC and RuBPCO were 103 and 86 kJ mol−1, respectively, for H. undatus and 77 and 49 kJ mol−1, respectively, for S. megalanthus, within the range determined for a diverse group of species using different methodologies. Above 28 °C, P N decreased an average of 58 % per °C in Phase I and 30 % per °C in Phase II for the two species; such steep declines with temperature indicate that irrigation then may lead to only small enhancements in net CO2 uptake ability.  相似文献   

17.
X. K. Yuan 《Photosynthetica》2016,54(3):475-477
In order to investigate the effect of day/night temperature difference (DIF) on photosynthetic characteristics of tomato plants (Solanum lycopersicum, cv. Jinguan 5) at fruit stage, an experiment was carried out in climate chambers. Five day/night temperature regimes (16/34, 19/31, 25/25, 31/19, and 34/16°C) with respective DIFs of -18, -12, 0, +12, and +18 were used and measured at mean daily temperature of 25°C. The results showed that chlorophyll (Chl) a, Chl b, net photosynthetic rate (PN), stomatal conductance (gs), maximum quantum yield of PSII photochemistry (Fv/Fm), effective quantum yield of PSII photochemistry (?PSII), and photochemical quenching (qp) significantly increased under positive DIF, while they decreased with negative DIF. In contrast, the Chl a/b ratio and nonphotochemical quenching (NPQ) decreased under positive DIF, while increased with negative DIF. Chl a, Chl b, PN, gs, Fv/Fm, ?PSII, and qp were larger under +12 DIF than those at +18 DIF, while Chl a/b and NPQ showed an opposite trend.  相似文献   

18.
  • 1.1. The diffusional water permeability (Pd) of rabbit red blood cell (RBC) membrane has been monitored by a doping nuclear magnetic resonance (NMR) technique on control cells and following inhibition with p-chloromercuribenzene sulfonate (PCMBS).
  • 2.2. The values of Pd were around 6.3 × 10−3 cm/sec at 15°C, 7.0 × 10−3cm/sec at 20°C, 8.0 × 10−3 cm/sec at 25°C, 9.1 × 10−3 cm/sec at 30°C and10.7 × 10−3 cm/sec at 37°C.
  • 3.3. Systematic studies on the effects of PCMBS on water diffusion indicated that the maximal inhibition was reached in 15 min at 37°C with 0.5 mM PCMBS.
  • 4.4. The values of maximal inhibition were around 71–74% at all temperatures.
  • 5.5. The basal permeability to water was estimated as 1.6 × 10−3cm/sec at 15°C, 2.0 × 10−3cm/sec at 20°C, 2.4 × 10−3cm/sec at 25°C, 2.6 × 10−3cm/sec at 30°C, and 3.1× 10−3 cm/secat 37°C.
  • 6.6. The activation energy of water diffusion was around 18 kJ/mol and increased to 27 kcal/mol after incubation with PCMBS in conditions of maximal inhibition of water diffusion.
  • 7.7. The membrane polypeptide electrophoretic pattern of rabbit RBCs has been compared with its human counterpart.
  • 8.8. The rabbit membrane contained a higher amount of spectrin (bands 1 and 2), while the band 6 (glyceraldehyde-3-phosphate dehydrogenase) was markedly less intense.
  • 9.9. Considerable differences in the electrophoretic patterns of the two sources of RBC membranes appeared in the bands migrating in the band 4.5 region and in front of band 7, where some polypeptides were apparent in higher amounts in the rabbit RBC membrane.
  相似文献   

19.
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   

20.
Two series of experiments were carried out to determine the relation of the rate of phosphorus and nitrogen excretion by the planktonic rotifers to ambient temperature and individual body weights of these animals. The following formulas describing this relation were obtained: EP=0.0154 W?1.27 e0.096T EN=0.0879 W?1.01 e0.088 T, where EP and EN denote the rate of P and N excretion, respectively, in µg · mg dry wt?1 · h?1, W is body weight in µg dry weight, and T is temperature in °C.  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