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.
R. H. Cherry 《BioControl》1979,24(1):35-39
Seasonally acclimatized citrus blackfly (CBF),Aleurocanthus woglumi Ashby, and its parasite,Amitus hesperidum Silv. were exposed to extreme temperatures for 3-hour periods and survivorship measured. In summer experiments, the LD50 for CBF adults and 1st instar nymphs occurred between 40° and 45°C. The LD50 forA. hesperidum adults occurred between 35° and 40°C and increased with decreased humidity, although some adults could continue to emerge from parasitized CBF up to 45°C. In winter experiments, the LD50 for CBF eggs and adults occurred between ?5° and ?10°C and for other CBF stages between ?10° and ?15°C. The LD50 for adultA. hesperidum, and survival and emergence of the parasites from CBF both occurred between ?10° and ?15°C. Comparing data from the 2 seasons shows that adults are the most temperature sensitive stage of CBF and the parasite is more temperature sensitive than CBF, especially at higher temperatures. Correlating lethal temperature data with field meteorological conditions shows that short-term temperature exposures cannot be expected to stop the potential spread of CBF orA. hesperidum through Florida.  相似文献   

2.
Suspending erythrocytes in medium containing sucrose prevented heat-induced lysis at its early stage. This allowed determination of the thermohaemolysis-related ion permeability by measuring the initial rate of the stipulated shrinkage of erythrocytes. Thus, correspondingly, the coefficient P of the ion permeability was calculated for heated human erythrocytes using ouabain-pretreated cells in 37–45°C range and intact cells in 50–58°C range. The values obtained for P obeyed a straight line Arrhenius plot over the entire 37–58°C range suggesting that the ion permeability was activated by a single mechanism earlier identified as relevant to thermohaemolysis. At the 37–58°C range, the activation energy of the P was 250±15 kJ/mol which was markedly different from the value of 56 kJ/mol known for the 10–37°C range. For erythrocytes from five mammals, similar temperature dependencies of the P were obtained over 45–60°C range. For erythrocytes from all species, excluding horse, the P, extrapolated at 37°C, had a value comparable with the known coefficient of the passive, ouabain-insensitive cation permeability at 37°C. For ouabain-treated human erythrocytes at 37°C, the period of thermohaemolysis-related shrinkage in sucrose containing media was found to be about six times shorter than the life-span of intact cells which substantiated the role of the active transport in balancing the thermohaemolysis-related diffusion of ions at 37°C. Consequently, the thermal resistance of erythrocytes, which was earlier related to their sphingomyelin content, was now found also to be in good correlation with their life-span in the circulation of 11 mammals.  相似文献   

3.
  • 1.1. Brain (hypothalamic), skin and body temperatures were measured in hand-reared acclimated (Acc, n = 5) and non-acclimated (NAcc, n =7) rock pigeons (Columba livia, mean body mass 237 g) exposed to increasing ambient temperatures (Ta) (30–60°C) and low humidities.
  • 2.2. In non-panting Acc birds, brain temperature gradually increased from 40.1 ± 0.4°C at 30°C to 41.2 ± 0.4°C at 60°C Ta. A mean body temperature (Tb) of 41.2 ± 0.2°C was measured at Ta up to 50°C; an increase of 1.1°C was observed at 60°C (Tb 42.2 ±0.6°C).
  • 3.3. In Acc panting birds exposed for 2 hr to 60°C, Thy was 41.9 ± 0.8°C and Ts was somewhat (but insignificantly) higher, i.e., 42.2 ± 0.7°C. It looks as if both values were increased as a result of a slight hyperthermia that developed (Tb = 43.5 ± 0.9°C).
  • 4.4. The significance of the present results for evaluating neuronal thermoresponsiveness of birds' hypothalamus is discussed.
  相似文献   

4.
Abstract. Factors underlying the process of photosynthetic acclimation to temperature were investigated for the shrub Nerium oleander L. Ramets of a single clone were grown under day/night temperature regimes of 20°C/15°C or 45°C/32°C. Plants grown at the lower temperature regime possessed rates of photosynthesis twice that of the high-temperature grown plants when CO2 fixation was measured at 20°C. In contrast, the plants grown at the high-temperature regime had twice the rate of CO2 fixation of the 20°C/l 5°C-grown plants at a measurement temperature of 45° C. It was determined that the ability to acclimate to changes in temperature regime was present in fully mature leaves. A reciprocal transfer of plants between the two growth regimes resulted in the appearance of the CO2 fixation characteristics appropriate to the new growth temperature after 12–14d. The response of CO2 fixation to light, temperature, and CO2 partial pressure and the temperature responses of soluble and membrane-bound photosynthetic enzyme systems were analysed to determine which components might be responsible for the superior photosynthetic performance of the 20°C/I5°C-grown plants at 20°C, and the enhanced high-temperature stability of the 45°C/32°C plants. The measured photosynthetic capacity of the 20°C/15°C plants could not be attributed to gross morphological, stomatal, or other physical changes, or to a general increase in the concentration of components of the photosynthetic process. Only a single enzyme, Fru-P2 phosphatase, was affected to an extent similar to that of photosynthesis. The enhanced thermal stability of the 45°C/32°C plants may be attributed primarily to an enhanced stability of the chloroplast membrane-bound enzymatic activities and the stability of the photosynthetic carbon metabolism enzymes which require lighl for activation.  相似文献   

5.
Yordanov  I.  Velikova  V.  Tsonev  T. 《Photosynthetica》1999,37(3):447-457
Fifteen-day-old bean plants (Phaseolus vulgaris L.) grown in a climatic chamber were exposed to water deficit (WD) and high temperature (HT) stresses applied separately or in combination. Changes in chlorophyll fluorescence quenching were investigated. Bean plants that endured mild (42 °C, 5 h for 2 d) WD separately or in combination with HT did not change their qP and qN quenching (measured at 25 °C) compared with those of the control. After 5 min testing at 45 °C, qP in control and droughted plants strongly decreased, while qP of plants that experienced combined WD+HT stress was insignificantly influenced, suggesting the acclimation effect of HT treatments. At more severe stresses (after 3 d-treatment), qP measured at 25 °C was the lowest in WD+HT plants and qN values were the highest. But when measured at 45 °C, qP of WD+HT plants had practically the same values as at 25 °C. Under these conditions qP of WD plants also showed an adaptation to HT. Twenty-four hours after recovery, the unfavourable effects of the stresses were strongly reduced when measured at 25 °C, but they were still present when measured at 45 °C. Positive effect of the carbamide cytokinin 4-PU-30 was well expressed only in droughted plants. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

6.
The compositions of glycolipids in the following seven strains of green photosynthetic bacteria were investigated at the molecular level using LC–MS coupled with an evaporative light scattering detector: Chlorobium (Chl.) limicola strains Larsen (30 °C as the optimal cultivation temperature) and DSM245 (30 °C), Chlorobaculum (Cba.) tepidum strain ATCC49652 (45 °C), Cba. parvum strain NCIB8327 (30 °C), Cba. limnaeum strain 1549 (30 °C), Chl. phaeovibrioides DSM269 (30 °C), and Chloroflexus (Cfl.) aurantiacus strain J-10-fl (55 °C). Dependence of the molecular structures of glycolipids including the chain-length of their acyl groups upon bacterial cultivation temperatures was clearly observed. The organisms with their optimal temperatures of 30, 45, and 55 °C dominantly accumulated glycolipids possessing the acyl chains in the range of C15–C16, C16–C17, and C18–C20, respectively. Cba. tepidum with an optimal temperature of 45 °C preferred the insertion of a methylene group to produce finally a C17-cyclopropane chain. Cfl. aurantiacus cultured optimally at 55 °C caused a drastic increase in the chain-length. Notably, the length of such acyl groups corresponded to that of the esterifying chain in the 17-propionate residues of self-aggregative bacteriochlorophylls-c/d/e, indicating stabilization of their supramolecular structures through hydrophobic interactions among those hydrocarbon chains. Based on the detailed compositions of glycolipids, a survival strategy of green photosynthetic bacteria grown in the wide range of temperatures is discussed.  相似文献   

7.
Restrictions to photosynthesis can limit plant growth at high temperature in a variety of ways. In addition to increasing photorespiration, moderately high temperatures (35–42 °C) can cause direct injury to the photosynthetic apparatus. Both carbon metabolism and thylakoid reactions have been suggested as the primary site of injury at these temperatures. In the present study this issue was addressed by first characterizing leaf temperature dynamics in Pima cotton (Gossypium barbadense) grown under irrigation in the US desert south‐west. It was found that cotton leaves repeatedly reached temperatures above 40 °C and could fluctuate as much as 8 or 10 °C in a matter of seconds. Laboratory studies revealed a maximum photosynthetic rate at 30–33 °C that declined by 22% at 45 °C. The majority of the inhibition persisted upon return to 30 °C. The mechanism of this limitation was assessed by measuring the response of photosynthesis to CO2 in the laboratory. The first time a cotton leaf (grown at 30 °C) was exposed to 45 °C, photosynthetic electron transport was stimulated (at high CO2) because of an increased flux through the photorespiratory pathway. However, upon cooling back to 30 °C, photosynthetic electron transport was inhibited and fell substantially below the level measured before the heat treatment. In the field, the response of assimilation (A) to various internal levels of CO2 (Ci) revealed that photosynthesis was limited by ribulose‐1,5‐bisphosphate (RuBP) regeneration at normal levels of CO2 (presumably because of limitations in thylakoid reactions needed to support RuBP regeneration). There was no evidence of a ribulose‐1,5‐bisphosphate carboxylase/oxygenase (Rubisco) limitation at air levels of CO2 and at no point on any of 30 ACi curves measured on leaves at temperatures from 28 to 39 °C was RuBP regeneration capacity measured to be in substantial excess of the capacity of Rubisco to use RuBP. It is therefore concluded that photosynthesis in field‐grown Pima cotton leaves is functionally limited by photosynthetic electron transport and RuBP regeneration capacity, not Rubisco activity.  相似文献   

8.
《Plant science》1987,49(2):75-79
The photosynthetic activity of leaf slices from Spinacia oleracea L., Cucumis sativus L. and Nerium oleander L. was measured in 25° C immediately after preincubation for 2.5 h at various photon flux densities (PFD) with chilling at 4°C, or at a moderate (450 μmol m−2 s−1) PFD with various temperatures below 25°C. Inhibition of photosynthesis was evident in C. sativus and 45°C-grown N. oleander after preincubation at 4°C at all PFD. The inhibition was most severe at fluxes in excess of the moderate PFD under which the plants were grown. Photosynthesis in S. oleracea and 20°C-grown N. oleander was not inhibited at 4°C unless the PFD was in excess of this moderate PFD. The inhibition of photosynthesis was initiated in C. sativus below 13°C, and in 45°C-grown N. oleander below 8°C. A phase transition in the polar lipids from the thylakoids of these plants was detected at about the same temperatures. For S. oleracea and 20°C-grown N. oleander preincubated under the same conditions, there was no inhibition of photosynthesis and no phase transition above 0°C. These results show that some component of photosynthesis was disrupted in the light at temperatures below that of the phase transition in the thylakoid polar lipids.  相似文献   

9.
Three trypsins (TRY-ES) were purified from Antarctic krill (Euphausia superba) by ammonium sulfate precipitation, ion-exchange and gel-filtration chromatography, with relative molecular mass of 28.7, 28.8 and 29.2 kDa respectively. The TRY-ES was inhibited by specific trypsin inhibitors (benzamidine, STI, CHOM and TLCK), with optimum temperature at 40 (Trypsin I), 45 (Trypsin II) and 40 °C (Trypsin III) repetitively. The TRY-ES was stabled between 5 and 40 °C, which was consistent with the red shift in fluorescence intensity peak at 40 °C (Trypsin I) and 45 °C (Trypsin II and Trypsin III) and blue shift at 40 °C (Trypsin II and Trypsin III). The K cat/K m values of the TRY-ES was 14.28, 9.46 and 5.93 mM?1s?1 respectively, 1.1–10.2 folds higher than trypsins from other crustacean and mammal, which was supported by the differences in thermodynamics parameters, the free energy, enthalpy, and entropy of benzamidine and the TRY-ES system.  相似文献   

10.
《Fungal biology》2020,124(6):571-578
Botryosphaeriaceae fungi are phytopathogens and human opportunists. The influence of temperature on the phytotoxicity and cytotoxicity of culture filtrates of five Botryosphaeriaceae species was investigated. All culture filtrates of fungi grown at 25 °C were phytotoxic: symptoms were evaluated based on visual inspection of necrosis areas and on the maximum quantum yield of photosystem II, Fv/Fm. Diplodia corticola and Neofusicoccum kwambonambiense were the most phytotoxic, followed by Neofusicoccum parvum CAA704 and Botryosphaeria dothidea. Phytotoxicity dramatically decreased when strains were grown at 37 °C, except for B. dothidea. All strains, except N. parvum CAA366 and Neofusicoccum eucalyptorum, grown either at 25 °C or 37 °C, were toxic to mammalian cells; at 25 °C and at 37°C, D. corticola and B. dothidea were the most cytotoxic, respectively. Although the toxicity of B. dothidea to both cell lines and of N. kwambonambiense to Vero cells increased with temperature, the opposite was found for the other species tested. Our results suggest that temperature modulates the expression of toxic compounds that, in a scenario of a global increase of temperature, may contribute to new plant infections but also human infections, especially in the case of B. dothidea.  相似文献   

11.
  • 1.1. The ambient temperature of embryos of pipped eggs was reduced from 38 to 28°C for a period of 45 min.
  • 2.2. The blood PCO2 was lower and the blood more alkaline at 28°C than at 38°C.
  • 3.3. At 28°C plasma [HCO3] ] was lower than predicted from the blood buffer line determined in vitro.
  • 4.4. The plasma concentrations of strong ions and lactate were the same at both temperatures.
  • 5.5. After the ambient temperature had been returned to 38°C for a period of 45 min, blood pH was more acidic than before cooling, but there was no difference in blood PCO2.
  • 6.6. The plasma [HCO3] was the same as that at 28°C and plasma [K+] was higher than before cooling.
  • 7.7. The results arc discussed in relation to the factors affecting blood pH in embryos at this stage of development.
  相似文献   

12.
Ventilation was measured directly in the hagfish, Myxine glutinosa L., by means of an electro-magnetic blood flowmeter. Ventilatory flow and frequency increased from 0.86 ± 0.27 ml·min?, and 18.2 ± 5.1·min?, respectively, at 7°C to 1.70 ± 0.20 ml·min?, and 70.1 ± 9.5·min? at 15 ·C.Standard oxygen consumption,V?O2, was measured in non-buried hagfish. V?O2 was 0.57 ± 0.17μl O2·g?1·min?1 at 7°C, and 0.85 ± 0.12μl O2·g?1·min?1 at 15°C.  相似文献   

13.
Aspergillus terreus NTOU4989 was isolated from sulfur sediment collected at Kueishan Island Hydrothermal Vent Field, Taiwan, and was shown to be able to grow at 45 °C and pH 3 in seawater, characteristics of hydrothermal vent sites. Fungi are one of the major decomposers in the marine environment, and to find out whether A. terreus NTOU4989 can fulfill such a role at the site, this study investigated its metabolic activity using the Biolog FF MicroPlate™ under different growth conditions (45 °C/pH 3, 45 °C/pH 7, 25 °C/pH 3, 25 °C/pH 7) with/without addition of Al3+, Fe2+ and Mn2+, major metal ions in the vent effluent. The greatest metabolic activity was observed at 25 °C/pH 3 with/without the metal ions based on substrate richness (RS, metabolic richness), average well colour development (AWCD, metabolic capacity) and average well turbidity development (AWTD, mycelial production). Significant low RS, AWCD and AWTD were observed at 45 °C, suggesting that A. terreus NTOU4989 is a thermotolerant fungus. The presence of the metal ions did not have a significant effect on its metabolic activity. Monomers of lignin, cellulose and hemicellulose, and common sugars and amino acids of algae were utilized in selected conditions, providing evidence of wood and algal decomposition by the fungus at the hydrothermal vent site.  相似文献   

14.
Thermal dependence of clearance rate (CR: l h?1), standard (SMR: J h?1) and routine metabolic rates (RMR: J h?1), were analyzed in fast (F)- and slow (S)-growing juveniles of the clam Ruditapes philippinarum. Physiological rates were measured at the maintenance temperature (17 °C), and compared with measurements performed at 10 and 24 °C after 16 h and 14 days to analyze acute and acclimated responses, respectively. Metabolic rates (both RMR and SMR) differed significantly between F and S seeds, irrespective of temperature. Mass-specific CRs were not different for F and S seeds but were significantly higher in F clams for rates standardized according to allometric size-scaling rules. Acute thermal dependency of CR was equal for F and S clams: mean Q 10 were ≈3 and 2 in temperature ranges of 10–17 and 17–24 °C, respectively. CR did not change after 2 weeks of acclimation to temperatures. Acute thermal effects on SMR were similar in both groups (Q 10 ≈ 1 and 1.6 in temperature ranges of 10–17 and 17–24 °C, respectively). Large differences between groups were found in the acute thermal dependence of RMR: Q 10 in F clams (≈1.2 and 1.9 at temperature ranges of 10–17 and 17–24 °C, respectively) were similar to those found for SMR (Q 10 = 1.0 and 1.7). In contrast, RMR of S clams exhibited maximum thermal dependence (Q 10 = 3.1) at 10–17 °C and become depressed at higher temperatures (Q 10 = 0.9 at 17–24 °C). A recovery of RMR in S clams was recorded upon acclimation to 24 °C. Contrasting metabolic patterns between fast and slow growers are interpreted as a consequence of differential thermal sensitivity of the fraction of metabolism associated to food processing and assimilation.  相似文献   

15.
In the present study, two genotypes each of maize and rice were compared for their response to varying degrees of temperature stress (35/30, 40/35, 45/40°C) with controls growing at 30/25°C. At elevated temperatures of 40/35 and 45/40°C, the rice genotypes were inhibited to a significantly higher extent, especially for their shoot growth compared to maize genotypes. The stress injury measured as damage to membranes, loss of chlorophyll and reduction in leaf water status was significantly higher in rice plants, especially at 45/40°C. The components of oxidative stress particularly the level of malondialdehyde was significantly greater in rice plants while the differences for hydrogen peroxide concentrations were small at 40/35 and 45/40°C. The expression of enzymatic antioxidants like catalase, ascorbate peroxidase and glutathione reductase was found to be higher in maize plants compared to rice plants while no variations existed for superoxide dismutase at 45/40°C. In addition, the non-enzymatic antioxidants like ascorbic acid, glutathione and proline were maintained at significantly greater levels at 45/40°C in maize than in rice genotypes. These findings suggested that maize genotypes were able to retain their growth under high-temperature conditions partly due to their superior ability to cope up with oxidative damage by heat stress compared to rice genotypes. Since, maize and rice belong to C4 and C3 plant groups, respectively, these observations may also reflect the relative sensitivity of these plant groups to heat stress.  相似文献   

16.
The kinetic properties of glucokinase (GLK) from the liver of active and hibernating ground squirrels Spermophilus undulatus have been studied. Entrance of ground squirrels into hibernation from their active state is accompanied by a sharp decrease in blood glucose (Glc) level (from 14 to 2.9 mM) and with a significant (7-fold) decrease of GLK activity in the liver cytoplasm. Preparations of native GLK practically devoid of other molecular forms of hexokinase were obtained from the liver of active and hibernating ground squirrels. The dependence of GLK activity upon Glc concentration for the enzyme from active ground squirrel liver showed a pronounced sigmoid character (Hill coefficient, h = 1.70 and S 0.5 = 6.23 mM; the experiments were conducted at 25°C in the presence of enzyme stabilizers, K+ and DTT). The same dependence of enzyme activity on Glc concentration was found for GLK from rat liver. However, on decreasing the temperature to 2°C (simulation of hibernation conditions), this dependency became almost hyperbolic (h = 1.16) and GLK affinity for substrate was reduced (S 0.5 = 23 mM). These parameters for hibernating ground squirrels (body temperature 5°C) at 25°C were found to be practically equal to the corresponding values obtained for GLK from the liver of active animals (h = 1.60, S 0.5 = 9.0 mM, respectively); at 2°C sigmoid character was less expressed and affinity for Glc was drastically decreased (h = 1.20, S 0.5 = 45 mM). The calculations of GLK activity in the liver of hibernating ground squirrels based on enzyme kinetic characteristics and seasonal changes in blood Glc concentrations have shown that GLK activity in the liver of hibernating ground squirrels is decreased about 5500-fold.  相似文献   

17.
Abstract: Growth in elevated CO2 led to an increase in biomass production per plant as a result of enhanced carbon uptake and lower rates of respiration, compared to ambient CO2-grown plants. No down-regulation of photosynthesis was found after six months of growth under elevated CO2. Photosynthetic rates at 15°C or 35 °C were also higher in elevated than in ambient CO2-grown plants, when measured at their respective CO2 growth condition. Stomata of elevated CO2-grown plants were less responsive to temperature as compared to ambient CO2 plants. The after effect of a heat-shock treatment (4 h at 45 °C in a chamber with 80% of relative humidity and 800–1000 tmol m-2 s-1 photon flux density) on Amax was less in elevated than in ambient CO2-grown plants. At the photochemical level, the negative effect of the heat-shock treatment was slightly more pronounced in ambient than in elevated CO2-grown plants. A greater tolerance to oxidative stress caused by high temperatures in elevated CO2-grown plants, in comparison to ambient CO2 plants, is suggested by the increase in superoxide dismutase activity, after 1 h at 45 °C, as well as its relatively high activity after 2 and 4 h of the heat shock in the elevated CO2-grown plants in contrast with the decrease to residual levels of superoxide dismutase activity in ambient CO2-grown plants immediately after 1 h at 45 °C. The observed increase in catalase after 1 h at 45 °C in both ambient and elevated CO2-grown plants, can be ascribed to the higher rates of photorespiration and respiration under this high temperature.  相似文献   

18.
A β-1,3-glucanase with a molecular mass of 33 kDa was isolated in the homogeneous state from a crystalline stalk of the commercially available Vietnamese edible mussel Perna viridis. It hydrolyzes β-1,3-bonds in glucans and is capable of catalyzing the transglycosylation reaction. The β-1,3-glucanase has a K m value of 0.3 mg/ml for the hydrolysis of laminaran and shows a maximum activity in the pH range from 4 to 6.5 and at 45°C. Its half-inactivation time is 180 min at 45°C and 20 min at 50°C. The enzyme was ascribed to glucan-endo-(1 → 3)-β-D-glucosidases (EC 3.2.1.39). The enzyme could be used in the structure determination of β-1,3-glucans and enzymatic synthesis of new carbohydrate-containing compounds.  相似文献   

19.
20.
Short-term receptor regulation by agonists is a well-known phenomenon for a number of receptors, including β-adrenergic receptors, and has been associated with receptor changes revealed by radioligand binding. In the present study, we investigated the rapid changes in α1-adrenergic receptors induced by agonists. α1-receptors were studied on DDT1 MF-2 smooth muscle cells (DDT1-MF-2 cells) by specific [3H]prazosin binding. In competition binding on membranes and on intact cells at 4°C or at 37°C in 1-min assays, agonists competed for a single class of sites with relatively high affinity. By contrast, in equilibrium binding at 37°C on intact cells agonists competed with two receptor forms (high- and low-affinity). We quantified the receptors in the high-affinity form by measuring the [3H]prazosin binding inhibited by 20 μM norepinephrine (this concentration selectively saturated the high-affinity sites). The low-affinity sites were measured by subtracting the binding of [3H]prazosin to the high-affinity sites from the total specific binding. High-affinity receptors were 85% of the total sites in binding experiments at 4°C, but only 30% at 37°C. On DDT1-MF-2 cells preequilibrated with [3H]prazosin at 4°C, and then shifted to 37°C for a few minutes, norepinephrine selectively reduced the high-affinity sites by 30%. We suggest that at 4°C it is the native form of α1-receptors that is measured, with most of the sites in the high-affinity form, while during incubation at 37°C the norepinephrine present in the binding assay converts most of the receptors to an apparent low-affinity form, so that they are no longer recognized by 20 μM norepinephrine. The nature of this low-affinity form was further investigated. On DDT1-MF-2 cells preincubated with the agonist and then extensively washed at 4°C (to maintain the receptor changes induced by the agonist) the number of receptors recognized by [3H]prazosin at 4°C was reduced by 38%. After fragmentation of the cells, the number of receptors measured at 4°C was the same in control and norepinephrine-treated cells, suggesting that the disruption of cellular integrity might expose the receptors which are probably sequestered after agonist treatment. In conclusion, the appearance of the low affinity for agonists at 37°C may be due to the agonist-induced sequestration of α1-adrenergic receptors, resulting in a limited accessibility to hydrophilic ligands.  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