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.
Anionic (Yb) rat liver glutathione-S-transferases are susceptible to temperature or pH dependent transitions to more basic forms of this class of proteins. At elevated temperatures (25-30 degrees) or at pH values above 9.0 the protein is rapidly and irreversibly converted to forms that are no longer retained by anion exchange resins and display basic components in chromatofocusing systems, because bound glutathione is removed at the higher temperatures or pH. Sharp increases in enzymatic activity with 1,2-dichloro-4-nitrobenzene as a substrate, accompany the temperature induced changes. Microheterogeneity patterns for this protein are contingent upon these interconversions, and the results explain apparent variations in relative amounts of the multiple forms under different conditions in terms of glutathione binding.  相似文献   

2.
The solubility of human plasma fibronectin in concentrated ammonium sulfate solutions was measured at pH 7.0 and varying temperatures as well as at 25 degrees C and varying pHs. The salting-out parameters, KS and beta were found to increase linearly with temperature in the range 5 degrees-50 degrees C. KS-pH and beta-pH profiles were found to have maxima at pH 7.0. The dependence of both of the solubility parameters of plasma fibronectin on temperature and pH was thus found to be anomalous. The possibility of a correlation between the heparin-induced cryoprecipitation of fibronectin and the dependence of its solubility parameters on pH and temperature is considered. It is suggested that heparin-induced precipitation of human plasma fibronectin at low temperatures is caused by (i) a cold effect and (ii) conformational change in the protein due to heparin binding.  相似文献   

3.
Although higher temperatures strongly stimulate ectothermic metabolic rates, they only slightly increase oxygen diffusion rates and decrease oxygen solubility. Consequently, we predicted that insect gas exchange systems would have more difficulty meeting tissue oxygen demands at higher temperatures. In this study, Drosophila melanogaster were reared from egg to adult in hyperoxic (40%), hypoxic (10%), and normoxic (21%) conditions and in temperatures ranging from 15 degrees -31.5 degrees C to examine the interactive effect of temperature and oxygen on development. Hyperoxia generally increased mass and growth rate at higher rearing temperatures. At lower rearing temperatures, however, hyperoxia had a very small effect on mass, did not affect growth rate, and lengthened time to eclosion. Relative to normoxia, flies reared in hypoxic conditions were generally smaller (mass and thorax length), had longer eclosion times, slower growth rates, and reduced survival. At cooler temperatures, hypoxia had relatively modest or nonsignificant effects on development, while at higher temperatures, the effects of hypoxia were large. These results suggest that higher temperatures reduce oxygen delivery capacity relative to tissue oxygen needs, which may partially explain why ectotherms are smaller when development occurs at higher temperatures.  相似文献   

4.
31P NMR was applied to an examination of the freeze-tolerant larvae of the gall fly, Eurosta solidaginis. Resonances from sugar phosphates, inorganic phosphate, adenylates and arginine phosphate were identified. Two peaks of Pi were identified corresponding to intracellular and extracellular Pi. Anoxia produced an expected decrease in peak intensities of ATP and arginine phosphate while the peak of intracellular Pi was enhanced and shifted to indicate intracellular acidification during anoxia. Spectra of whole larvae were monitored over a temperature range from -30 degrees to +25 degrees C. No abrupt alterations in the spectra were seen at the point of extracellular freezing which occurs at about -8 degrees C but temperature had dramatic effects upon the peak intensities of ATP and arginine phosphate. A reversible increase/decrease in peak intensities, relative to Pi, was observed as temperature was raised/lowered. At 15 degrees and -20 degrees C, the beta peak of ATP was 64% and 2% of the peak intensity of Pi while that of arginine phosphate was 78% and 11%, respectively. This temperature effect was not an artifact of instrumentation (as model solutions containing Pi, ATP and arginine phosphate did not show this effect) or a result of changes in the total amounts of these compounds in the cell with temperature. Rather it is apparent that these molecules become restricted in their rotational movement as temperature is lowered perhaps via binding to subcellular components. Changes in the amounts of freely soluble ATP and arginine phosphate with temperature could have important implications for metabolism and its control. Analysis of the effect of temperature on the chemical shift of Pi was also used to determine pH in the intracellular and extracellular compartments. Temperature change had no effect on extracellular (hemolymph) pH which remained constant at 6.1-6.3. Intracellular pH varied with temperature, however, from pH 6.8 at 15 degrees C to pH 7.3 at -12 degrees C with a change, delta pH/delta 0, of -0.0185 degrees C consistent with alphastat regulation.  相似文献   

5.
Lipo-chitin oligosaccharides (LCOs) are usually produced and isolated for structural analysis from bacteria cultured under laboratory rather than field conditions. We have studied the influence of bacterial growth temperature on the LCO structures produced by different Rhizobium leguminosarum strains, using thin-layer chromatographic, high-performance liquid chromatographic, and mass spectrometric analyses. Wild-type R. leguminosarum bv. viciae A1 was shown to produce larger relative amounts of nodX-mediated, acetylated LCOs at 12 degrees C than at 28 degrees C, indicating that the activity of nodX (a gene encoding an LCO O-acetyl transferase) is temperature dependent. Interestingly, symbiotic resistance genes sym1 and sym2 found in primitive pea cultivars are also temperature sensitive, only being active at low temperatures, at which they block nodulation by R. leguminosarum bv. viciae strains lacking nodX. We therefore propose that the gene-for-gene relationship between plant and bacterium has a temperature-sensitive mechanism as an adaptation to environmental conditions. An R. leguminosarum bv. trifolii strain was also shown to produce larger relative amounts of nodX-mediated, acetylated LCOs at 12 degrees C than at 28 degrees C. The major components synthesized by the two strains are produced at both temperatures but in different relative amounts, while some minor components are only produced at one of the two temperatures.  相似文献   

6.
M G Mulkerrin  R Wetzel 《Biochemistry》1989,28(16):6556-6561
Heated at pH 6.0 and at 50 degrees C, human interferon gamma (HuIFN-gamma) is inactivated via the formation of insoluble aggregates. At pH 6.0, the aggregation rate increases with temperature from 40 to 65 degrees C. There is a temperature-dependent time lag to aggregate formation observed in the generation of light-scattering particles at pH 6.0, and this correlates with the fast phase observed in the kinetics of reversible thermal unfolding. In addition, the dependence of aggregation kinetics on temperature closely follows the reversible melting curve. These observations suggest that at pH 6.0 irreversible thermal denaturation and aggregation depend on partial or complete unfolding of the molecule. At pH 5.0, also at 50 degrees C, the molecule is stable to irreversible aggregation. In reversible unfolding in 0.25 M guanidine hydrochloride, the Tm for HuIFN-gamma increases from 30.5 degrees C at pH 4.75 to 41.8 degrees C at pH 6.25, in analogy to the behavior of other globular proteins. These observations suggest that the relative instability of HuIFN-gamma to irreversible denaturation via aggregation at pH 6.0 compared to pH 5.0 is not due to an increased stability toward unfolding at the lower pH. Alternatively, stability at pH 5.0 must be due either to the improved solution properties of the unfolded state or to the improved solubility/decreased kinetic lifetime of an unfolding intermediate. Aggregation of HuIFN-gamma at 50 degrees C is half-maximal at pH 5.7, suggesting that protonation of one or both of the histidine residues may be involved in this stabilization.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
We have examined the effects of the local anesthetic tetracaine on the orientational and dynamic properties of glycolipid model membranes. We elected to study the interactions of tetracaine with the pure glycolipid 1,2-di-O-tetradecyl-3-O-(beta-D-glucopyranosyl)-sn-glycerol (beta-DTGL) and a mixture of beta-DTGL (20 mol%) in dimyristoylphosphatidylcholine (DMPC) by deuterium NMR (2H-NMR) spectroscopy. 2H-NMR spectra of beta-DTGL have been measured as a function of temperature in the presence of both the charged (pH 5.5) and uncharged forms (pH 9.5) of tetracaine. The results indicate that the anesthetic induces the formation of non-lamellar phases. Specifically, the incorporation of uncharged tetracaine results in the formation of a hexagonal phase which is stable from 52 to 60 degrees C. At lower pH, the spectrum at 52 degrees C is very reminescent of that of the beta-glucolipid alone in a bilayer environment, while as the temperature is elevated to 60 degrees C, a transition from a spectrum indicative of axial symmetry to one due to nearly isotropic motion or symmetry occurs, which may result from the formation of a cubic phase. Although it leads to an alteration in the phase behavior, the presence of tetracaine does not induce large changes in the headgroup orientation of beta-DTGL. In contrast to the pure glycolipid situation, the interaction of tetracaine with beta-DTGL (20 mol%) in DMPC does not trigger the formation of non-lamellar phases, but leads to a slight reduction in molecular ordering. The presence of the charged form of the local anesthetic near the aqueous interface of the bilayer appears to induce a small change in the conformation about the C2-C3 bond of the glycerol backbone of beta-DTGL in the mixed lipid system. Thus, the major influence of the local anesthetic on glycolipids is a change in the stability of the lamellar phase, facilitating conversion to phases with hexagonal or isotropic environments for the lipid molecules.  相似文献   

8.
This study evaluated the ability of mutagenic antineoplastic agents to vaporize at room temperature (23 degrees C) and 37 degrees C. A bacterial mutagenicity assay was used to determine the mutagenicity of these agents in the vapor phase. Open plates of bacteria were exposed to varying amounts of drug solutions in sealed glass containers for 24h. The drug solutions were prepared as they would be for patient treatment and were tested at 0.25, 0.5 and 1.0 ml of each drug solution per 10 l of air. Following exposure, the plates exposed at 23 degrees C were incubated an additional 48 h at 37 degrees C to allow for expression of mutations. Those exposed at 37 degrees C were incubated for an additional 24h at 37 degrees C. Carmustine, cyclophosphamide, ifosfamide, thiotepa, and mustargen demonstrated vaporization at 37 degrees C. Carmustine and mustargen also demonstrated significant vaporization at 23 degrees C, while cyclophosphamide demonstrated a 50% increase in revertants at this temperature. In addition, sodium azide, a known mutagen used as a control was also mutagenic as a vapor at both temperatures. Doxorubicin, cisplatin, etoposide, 5-fluorouracil and mitomycin were not detected as vaporizing in this assay. The study found that vaporization of standard solutions of some antineoplastic agents is possible at room temperature and increases as the temperature increases. Therefore, vaporization of spilled antineoplastic agents may present an additional route of exposure to healthcare workers through inhalation.  相似文献   

9.
1. Comparison of erythrocyte osmotic fragility (EOF) between various ectotherms and endotherms was investigated at 5, 25, and 38 degrees C. 2. We hypothesized that ectotherms might possess erythrocytes whose osmotic fragility would be less affected by temperature than those of endotherms. 3. Ectotherm erythrocytes were much more osmotically resistant than those of endotherms. 4. The EOF of ectotherms and endotherms showed similar responses to temperature. 5. It does not appear that the osmotic fragility of erythrocytes from ectotherms in this study are adapted to be less affected by temperature than those of endotherms. The highly osmotic resistant erythrocytes of ectotherms may alleviate the need for further adaptation for osmotic resistance.  相似文献   

10.
Potentiometric titration of the zwitter-ion of 7-aminoacetoxycephalosporanic acid (7-ACA) was performed and the constants of its ionization were estimated. The minimum solubility of the 7-ACA zwitter-ion (20 degrees C, 0.1 M NaCl) was determined and the solubility curve of 7-ACA at wide pH ranges was calculated. Equilibrium of the cationic, zwitter-ionic and anionic forms of 7-ACA was estimated as dependent on pH.  相似文献   

11.
The phase transition temperature of 1,2-distearoylglycerophosphocholine is reduced in presence of equimolar amounts of 1-O-(1'-alkenyl)-glycerophosphoethanolamine (ethanolamine lysoplasmalogen) from 53.3 degrees C-54.1 degrees C to 44.0 degrees C-44.9 degrees C at different pH (4.0; 7.2; 9.0; 10.5). 1-Acyl-glycerophosphoethanolamine leads to a smaller reduction of the 1,2-distearoyl-glycerophosphocholine transition temperature: 45.0 degrees C-46.2 degrees C at the same pH-values. 1-Alkyl-glycerophosphoethanolamine (hydrogenated ethanolamine lysoplasmalogen) possesses a transition temperature, which is 3.3 degrees C-4.9 degrees C higher than the hydrogenated 1-acyl-glycerophosphoethanolamine at each pH investigated. At pH 9.0 and, more pronounced, at pH 10.5 we find a reduction of the transition temperature for both these substances, whereas their transition temperature is nearly unchanged at pH 4.0 and 7.2. Our results clearly show that the ether-bonding in the lysoderivative of plasmalogen is responsible for the closer packing compared to the 1-acyl-glycerophosphoethanolamine.  相似文献   

12.
By the dynamics of human oxyhemoglobin coagulation in the presence of phenyl mercury acetate in tris-AcOH buffer, pH 7.2 the number of moles of PhHg+ stechiometrically bound with protein at different temperatures was estimated. Within the temperature range 15-30 degrees C this value is constant--32-34 mole per 1 mole of HBO2-tetramer. Within the range 30-40 degrees C it rises to approximately 40. Coagulation of oxyhemoglobin modified with PhHg+ cation is reversible in contrast to HBO2 coagulation modified with uncharged PhHgCl.  相似文献   

13.
In Drosophila, like most ectotherms, development at low temperature reduces growth rate but increases final adult size. Cultures were shifted from 25 degrees C to low (16.5 degrees C) or to high (29 degrees C) temperature at regular intervals through larval and pupal stages, and the flies of both sexes showed an increase or decrease, respectively, in the size of thorax, wing and abdominal tergite. Size changes in the wing blade resulted from changes in the size of the epidermal cells (with only a small increase in cell number in males reared at low temperature). The temperature-shifts became less effective as they were made at successively later developmental stages, demonstrating a cumulative effect of temperature on adult size. The thorax and wing develop from the same imaginal disc, with most cell division occurring in larval stages, but they differ in timing of temperature sensitivity, which extends only to pupariation or into the late pupal stage, respectively. Growth of the adult abdomen occurs largely after pupariation but its size is temperature-sensitive through both larval and pupal stages. We discuss growth control in Drosophila and the likely effects of temperature on food assimilation, growth efficiency and allocation of nutrients to the production of different tissues.  相似文献   

14.
Urban physiology: city ants possess high heat tolerance   总被引:1,自引:0,他引:1  
Urbanization has caused regional increases in temperature that exceed those measured on a global scale, leading to urban heat islands as much as 12 degrees C hotter than their surroundings. Optimality models predict ectotherms in urban areas should tolerate heat better and cold worse than ectotherms in rural areas. We tested these predications by measuring heat and cold tolerances of leaf-cutter ants from South America's largest city (S?o Paulo, Brazil). Specifically, we compared thermal tolerances of ants from inside and outside of the city. Knock-down resistance and chill-coma recovery were used as indicators of heat and cold tolerances, respectively. Ants from within the city took 20% longer to lose mobility at 42 degrees C than ants from outside the city. Interestingly, greater heat tolerance came at no obvious expense of cold tolerance; hence, our observations only partially support current theory. Our results indicate that thermal tolerances of some organisms can respond to rapid changes in climate. Predictive models should account for acclimatory and evolutionary responses during climate change.  相似文献   

15.
The effect of temperature on the amount of uncoupling protein mRNA in rat brown adipose tissue was examined after 1 and 14 days of exposure to cold. The relative amounts after 1 day, compared with rats kept at a thermoneutral temperature of 28 degrees C, were 3.2 at 19 degrees C, 3.3 at 11 degrees C, and 2.1 at 3 degrees C. This suggests that in warm-acclimated rats, a maximal response to a cold stimulus in brown adipose tissue is reached by 19 degrees C. In contrast to these results, the relative amounts of uncoupling protein mRNA after 14 days of cold exposure, compared with rats left at 28 degrees C, were 1.2 at 19 degrees C, 1.9 at 11 degrees C, and 2.1 at 3 degrees C. Since it is known that the amount of uncoupling protein in cold-acclimated rats increases continuously with decrease in temperature, the amount of protein reflects the mRNA levels during later times but not the initial time of exposure to cold.  相似文献   

16.
The rates of the alkaline denaturation of the covalently closed, circular DNAs (form I) of the replicative forms (RF) of phages G4, phi X174, and fd, and of plasmid pBR322 and phage PM2 have been measured at 0 degrees C and some at higher temperatures. These rates are orders of magnitude slower than the denaturation of linear DNA because of the increased stability of the helix to deprotonation that results from the accumulating positive superhelicity during denaturation. Denaturation reactions were initiated by rapid, infrasonic mixing (Camien, M.N., and Warner, R.C. (1984) Anal. Biochem. 138, 329-334), and their progress was measured by analytical ultracentrifugal analysis for the amounts of form I and denatured (Id) DNA after neutralization of the alkaline reaction. The comparative rates of the five DNAs varied over a wide range; the fastest, G4-RF, denatured at 500-fold the rate of the slowest, fd-RF. The differences are accounted for by the interaction of positive superhelicity with the sequence-dependent regions of relative helix stability in the various DNAs. Renaturation rates of Id DNAs varied similarly for Ids prepared at 0 degrees C, but only a few-fold for Ids prepared at 50 degrees C. The rate of denaturation of G4-RF was determined over a wide range of NaOH and NaCl concentrations at 0 degrees C, and the pHm was determined as a function of ionic strength and temperature. The effects of ionic strength have been analyzed in an application of the Manning ion condensation-screening theory (Manning, G.S. (1978) Q. Rev. Biophys. 11, 179-246) which is shown to account for the large destablizing effect of salts on the helix. The pH region of transition at 50 degrees C from renaturation to denaturation was examined, and it was found that the maximum rate of renaturation occurred at a pH about 0.05 units below the pHm.  相似文献   

17.
The interactions of pH (5.0, 6.0, and 7.0), temperature (19, 28, and 37 degrees C), and atmosphere (aerobic versus anaerobic) with NaCl (0, 1, 2, 3, 4, and 5%) on the growth of Salmonella typhimurium ATCC 14028 in defined glucose-mineral salts culture medium were evaluated. Response surface methodology was used to develop equations describing the response of S. typhimurium to environmental changes. The response to an increasing concentration of NaCl at any temperature tested was nonlinear. The maximum growth was predicted to occur at an NaCl concentration of 0.5%, a temperature of 19 degrees C, and an initial pH of 7.0 under aerobic growth conditions. The relative amounts of aerobic growth at 19 degrees C, pH 7.0, and NaCl concentrations of 0, 0.5, 1, 2, 3, 4, and 5% were predicted to be 99.2, 100.0, 98.8, 90.2, 73.5, 48.6, and 15.6%, respectively. Anaerobic growth conditions repressed the amount of growth relative to that under aerobic conditions, and the effects of NaCl and pH were additive at low salt concentrations; however, at higher salt levels anaerobiosis provided protection against the effects of NaCl.  相似文献   

18.
The interactions of pH (5.0, 6.0, and 7.0), temperature (19, 28, and 37 degrees C), and atmosphere (aerobic versus anaerobic) with NaCl (0, 1, 2, 3, 4, and 5%) on the growth of Salmonella typhimurium ATCC 14028 in defined glucose-mineral salts culture medium were evaluated. Response surface methodology was used to develop equations describing the response of S. typhimurium to environmental changes. The response to an increasing concentration of NaCl at any temperature tested was nonlinear. The maximum growth was predicted to occur at an NaCl concentration of 0.5%, a temperature of 19 degrees C, and an initial pH of 7.0 under aerobic growth conditions. The relative amounts of aerobic growth at 19 degrees C, pH 7.0, and NaCl concentrations of 0, 0.5, 1, 2, 3, 4, and 5% were predicted to be 99.2, 100.0, 98.8, 90.2, 73.5, 48.6, and 15.6%, respectively. Anaerobic growth conditions repressed the amount of growth relative to that under aerobic conditions, and the effects of NaCl and pH were additive at low salt concentrations; however, at higher salt levels anaerobiosis provided protection against the effects of NaCl.  相似文献   

19.
The ideal free distribution (IFD) predicts that optimal foragers will select foraging patches to maximize food rewards and that groups of foragers should thus be distributed between food patches in proportion to the availability of food in those patches. Because many of the underlying mechanisms of foraging are temperature dependent in ectotherms, the distribution of ectothermic foragers between food patches may similarly depend on temperature because the difference in fitness rewards between these patches may change with temperature. We tested the hypothesis that the distribution of Common Gartersnakes (Thamnophis sirtalis) between food patches can be explained by an IFD, but that conformance to an IFD weakens as temperature departs from the optimal temperature because fitness rewards, interference competition and the number of individuals foraging are highest at the optimal temperature. First, we determined the optimal temperature for foraging. Second, we examined group foraging at three temperatures and three density treatments. Search time was optimized at 27°C, handling time at 29°C and digestion time at 32°C. Gartersnakes did not match an IFD at any temperature, but their distribution did change with temperature: snakes at 20°C and at 30°C selected both food patches equally, while snakes at 25°C selected the low food patch more at low density and the high food patch more at high density. Food consumption and competition increased with temperature, and handling time decreased with temperature. Temperature therefore had a strong impact on foraging, but did not affect the IFD. Future work should examine temperature‐dependent foraging in ectotherms that are known to match an IFD.  相似文献   

20.
M Auger  H C Jarrell  I C Smith 《Biochemistry》1988,27(13):4660-4667
The interactions of the local anesthetic tetracaine with multilamellar dispersions of 1,2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC) and cholesterol have been investigated by deuterium nuclear magnetic resonance of specifically deuteriated tetracaines, DMPC and cholesterol. Experiments were performed at pH 5.5, when the anesthetic is primarily charged, and at pH 9.5, when it is primarily uncharged. The partition coefficients of the anesthetic in the membrane have been measured at both pH values for phosphatidylcholine bilayers with and without cholesterol. The higher partition coefficients obtained at pH 9.5 reflect the hydrophobic interactions between the uncharged form of the anesthetic and the hydrocarbon region of the bilayer. The lower partition coefficients for the DMPC/cholesterol system at both pH values suggest that cholesterol, which increases the order of the lipid chains, decreases the solubility of tetracaine into the bilayer. For phosphatidylcholine bilayers, it has been proposed [Boulanger, Y., Schreier, S., & Smith, I. C. P. (1981) Biochemistry 20, 6824-6830] that the charged tetracaine at low pH is located mostly at the phospholipid headgroup level while the uncharged tetracaine intercalates more deeply into the bilayer. The present study suggests that the location of tetracaine in the cholesterol-containing system is different from that in pure phosphatidylcholine bilayers: the anesthetic sits higher in the membrane. An increase in temperature results in a deeper penetration of the anesthetic into the bilayer. Moreover, the incorporation of the anesthetic into DMPC bilayers with or without cholesterol results in a reduction of the lipid order parameters both in the plateau and in the tail regions of the acyl chains, this effect being greater with the charged form of the anesthetic.  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