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 453 毫秒
1.
We have demonstrated that prolonged hypoxia can induce both thermotolerance and the synthesis of heat shock proteins in HA-1 Chinese hamster ovary cells. This tolerance was transient in nature: upon reaeration at 37 °C, HA-1 cells regained their “normal” heat response within 34 h.  相似文献   

2.
The influence of the growth medium and the growth temperature on the heat resistance of Citrobacter freundii has been established. Logarithmic growth phase cells grown on rich media have a higher heat resistance than cells of the same phase grown on minimal media. This finding was independent of type of carbon source in the growth medium, but the kind of carbon source has a definite influence on the heat resistance. Logarithmic phase cells grown at 37°C are much more heat stable than cells grown at 20 or 41°C. Stationary growth phase cells are much more heat resistant than logarithmic phase cells, whereas Mg2+-or glucose-starved cells are even slightly more heat stable than stationary phase cells.  相似文献   

3.
In this study, we examined the effect of concurrent low concentrations of sodium arsenite and mild heat shock temperatures on hsp30 and hsp70 gene expression in Xenopus A6 kidney epithelial cells. RNA blot hybridization and immunoblot analysis revealed that exposure of A6 cells to 1–10 µM sodium arsenite at a mild heat shock temperature of 30 °C enhanced hsp30 and hsp70 gene expression to a much greater extent than found with either stress individually. In cells treated simultaneously with 10 µM sodium arsenite and different heat shock temperatures, enhanced accumulation of HSP30 and HSP70 protein was first detected at 26 °C with larger responses at 28 and 30 °C. HSF1 activity was involved in combined stress-induced hsp gene expression since the HSF1 activation inhibitor, KNK437, inhibited HSP30 and HSP70 accumulation. Immunocytochemical analysis revealed that HSP30 was present in a granular pattern primarily in the cytoplasm in cells treated simultaneously with both stresses. Finally, prior exposure of A6 cells to concurrent sodium arsenite (10 µM) and heat shock (30 °C) treatment conferred thermotolerance since it protected them against a subsequent thermal challenge (37 °C). Acquired thermotolerance was not observed with cells treated with the two mild stresses individually.  相似文献   

4.
Tolerance of photosynthesis to high temperature in desert plants   总被引:9,自引:5,他引:4       下载免费PDF全文
Winter- and summertime-active desert annual species were grown at different temperatures to assess their capacity for photosynthetic acclimation. Thermal stability of photosynthesis was determined from responses of chlorophyll fluorescence to increased temperature. Photosynthesis in winter ephemerals grown at 28°C/21°C became unstable close to 41°C in contrast to the summer annuals which were stable up to about 46°C. Growth at higher temperature (43°C/32°C) resulted in increases in thermal stability of 5 to 7°C for the winter annuals and 3 to 4°C for the summer annuals, showing that temperature can provide the primary stimulus for acclimation of the photosynthetic apparatus. The magnitude of these changes was very similar to the range of field values observed for the respective floras, indicating that the thermal acclimation response under field conditions was qualitatively similar to that occurring under controlled growth conditions. Perennial species, co-existing with these annuals in the desert, were on average more thermostable. The cacti were exceptionally heat stable, the threshold for fluorescence increase averaging 55°C.  相似文献   

5.
Random and synchronous V79 cells were shifted from 37.5 °C to temperatures between 29 ° and 41 °C. Intermitotic time determinations of random cultures showed an increase in generation time and a broadening in the distribution of generation times in cells whose cycle spanned the temperature shift, but only a slight increase in generation time after one generation at temperatures between 34 °–40 °C. At 33.5 °C and below there was a stepwise increase in generation time. When cells grown at non-standard temperatures were allowed to habituate for 48 h at the altered temperature prior to analysis, the increase in median intermitotic time was slightly less in comparison to analyses done after only one generation following the temperature step. The Q10 for cell division of cells growing at temperatures from 34 ° to 40 °C was between 1.15 and 1.26, suggesting that the mammalian cell cycle is temperature compensated over a limited (6–7 °C) temperature span. Mammalian cells in culture appear to have the same capacity for temperature compensation in their cell cycle as do unicellular eukaryotes. The fact that cycle time at lower temperatures increases in a discrete manner is taken as evidence for a quantal clock.  相似文献   

6.
A laboratory-induced mutant with heat-sensitive development of the phagocytotic organelle has been isolated in Tetrahymena pyriformis, syngen 1; the mutant cells form food vacuoles at 30 °C, but not after incubation at 37 °C. Mutant cells transferred to 37 °C undergo a maximum of 3–5 doublings, but a sizeable fraction remains viable for several days. Results of temperature shift-up experiments reveal that an oral apparatus (OA) constructed at 30 °C remains functional at 37 °C, while one constructed at 37 °C is non-functional with regard to phagocytosis. Preliminary cytological observations reveal severe structural abnormalities of the OA. Thus the mutant appears to be primarily affected in the morphogenesis of the OA. The phenotypic effect of the mutation is reversible by a temperature shiftdown. Changes in phenotype caused by temperature shifts in either direction can occur even in stationary or starved cultures. Cell division is not required for the resumption of phagocytosis after a temperature shiftdown. Null-formers obtained at the first doubling after a temperature shift-up can divide at least once more, indicating that a functional OA is not required for cell division at any stage of the cell cycle. Mutants defective in phagocytosis may prove useful in gaining deeper understanding of this mechanism and its relationship to other cellular processes.  相似文献   

7.
Baudoinia compniacensis is a microfungus recently described as the principal agent of fouling known as “warehouse staining”, affecting building exteriors, fixtures and vegetation surfaces in areas proximate to distillery aging warehouses, commercial bakeries and other areas subject to low-level ethanol vapour exposure. The surfaces most affected tend to be highly exposed and undergo extreme diurnal temperature fluctuations. In previous work, we have demonstrated the existence of heat-inducible putative chaperone proteins that may also be induced by low-level exposures to ethanol vapour (e.g., <10 ppm). The present study investigated the cellular accumulation of trehalose, a disaccharide identified in some microorganisms to be important in the protection of cell components during adverse stress conditions, such as thermal stress. Following heat shock at 45 °C, we observed a 2.5-fold accumulation of trehalose relative to unheated controls maintained at 26 °C. Peak trehalose concentrations of 10 mg g−1 dry wt were seen at 90 min after heat treatment, followed by a gradual return to post-treatment by 150 min. Exposure of B. compniacensis cells to ethanol resulted in a similar increased accumulation of trehalose compared to unexposed controls. These findings imply that trehalose may be important in the tolerance of this fungus to abiotic stresses, such as heat and solvent exposure, and suggest future research directions for the control and prevention of warehouse staining.  相似文献   

8.
The membrane composition of Zymomonas mobilis changed dramatically in response to growth temperature. With increasing temperature, the proportion of vaccenic acid declined with an increase in myristic acid, the proportion of phosphatidylcholine and cardiolipin increased with decreases in phosphatidylethanolamine and phosphatidylglycerol, and the phospholipid/protein ratio of the membrane declined. These changes in membrane composition were correlated with changes in thermal tolerance and with changes in membrane fluidity. Cells grown at 20°C were more sensitive to inactivation at 45°C than were cells grown at 30°C, as expected. However, cells grown at 41°C (near the maximal growth temperature for Z. mobilis) were hypersensitive to thermal inactivation, suggesting that cells may be damaged during growth at this temperature. When cells were held at 45°C, soluble proteins from cells grown at 41°C were rapidly lost into the surrounding buffer in contrast to cells grown at lower temperatures. The synthesis of phospholipid-deficient membranes during growth at 41°C was proposed as being responsible for this increased thermal sensitivity.  相似文献   

9.
The interactions between concanvalin A and chick embryo fibroblasts, normal and infected with Rous sarcoma virus (RSV-BH) or its thermosensitive mutant RSV-BH-Ta, have been studied. Normal chick embryo cells and RSV-BH transformed cells showed at 4 and 25 °C a similar number of concanavalin A receptors per cell. Analysis of the binding data by the Scatchard relation showed that apparent changes in binding as a function of temperature are due to the thermodynamic properties of the process and and not to endocytosis. The lectin receptors on the cell surface of normal and RSV-BH infected cells showed homogeneity in their binding properties. Chick cells infected with RSV-BH-Ta showed a lectin binding behavior that was dependent on the temperature at which the cells were grown. At the permissive temperature for transformation (37 °C), the binding process was similar to that observed for normal and RSV-BH infected cells. At the nonpermissive temperature (41 °C), the cells showed at least two sets of concanavalin A receptors. The new set of receptors on the cell surface had a lower lectin affinity than those observed in the same cells at 37 °C.Chick cells infected with RSV-BH showed an enhanced agglutinability by concanavalin A, as compared with normal cells. Cells infected with RSV-BH-Ta showed a reversal of the correlation between increased concanavalin A agglutinability and the transformed state. At the permissive temperature for transformation, the cells were not agglutinable, whereas at the nonpermissive temperature they presented agglutinability indexes as high as those observed with RSV-BH infected cells. This enhanced agglutinability observed with cells maintained at the nonpermissive temperature for transformation may be related to the new set of low affinity receptors present at 41 °C.  相似文献   

10.
Membrane fatty acid composition and thermal resistance (D value) of Pediococcus sp. were determined for mid-exponential-phase (ME) and stationary-phase (ST) cells grown in tryptic soy broth (TSB) and tryptone-glucose-yeast extract (TGY) at 28 and 37°C. As the cells entered the stationary phase of growth, the unsaturated fatty acid, C18:1n11c, produced during the exponential phase of growth was converted to its cyclic form, C19:0Δ9c. This shift in membrane fatty acid composition was accompanied by an increase in the D values of this bacterium. Data from this study suggest that the membrane fatty acid composition of Pediococcus sp. is dependent on the growth conditions and that membrane fatty acid composition plays a critical role in thermal resistance. Thermal inactivation curves of Pediococcus sp. cells grown in TGY at 28°C indicated the presence of a cell population that is heterogeneous in thermal resistance. The growth of this bacterium in TGY at 37°C and in TSB at 28 and 37°C resulted in cell populations that were uniform in thermal resistance with a lag time for thermal inactivation. Thermal inactivation curves of ME and ST cultures were similar. The data presented here suggest that the cell population’s uniformity of thermal inactivation is independent of the growth phase of the culture.  相似文献   

11.
The influence of the physical state of the membrane on the swimming behaviour of Tetrahymena pyriformis was studied in cells with lipid-modified membranes. When the growth temperature of Tetrahymena cells was increased from 15°C to 34°C or decreased from 39°C to 15°C, their swimming velocity changed gradually in a similar to the adaptive change in membrane lipid composition. Therefore, such adaptive changes in swimming velocity were not observed during short exposures to a different environment. Tetrahymena cells adapted to 34°C swam at 570 μm/s. On incubation at 15°C these cells swam at 100 μm/s. When the temperature was increased to 34°C after a 90-min incubation at 15°C, the initial velocity was immediately recovered. On replacement of tetrahymanol with ergosterol, the swimming velocity of 34°C-grown cells decreased to 210 μm/s, and the cells ceased to move when the temperature was decreased to 15°C. To investigate the influence of the physical state of the membrane on the swimming velocity, total phospholipids were prepared from Tetrahymena cells grown under these different conditions. The fluidities of liposomes of these phospholipid were measured using stearate spin probe. The membrane fluidity of the cells cooled to 15°C increased gradually during incubation at 15°C. On the other hand, the fluidity of the heated cell decreased during incubation at 34°C. Replacement of tetrahymanol with ergosterol decreased the membrane fluidity markedly. Consequently, a good correlation was observed between swimming velocity and membrane fluidity; as the membrane fluidity increased, the swimming velocity increased linearly up to 600 μm/s. These results provide evidence for the regulation of the swimming behaviour by physical properties of the membrane.  相似文献   

12.
An accumulation in cells of unfolded proteins is believed to be the common signal triggering the induction of heat shock proteins (hsps). Accordingly, in Saccharomyces cerevisiae, inhibition of protein breakdown at 30°C with the proteasome inhibitor MG132 caused a coordinate induction of many heat shock proteins within 1 to 2 h. Concomitantly, MG132, at concentrations that had little or no effect on growth rate, caused a dramatic increase in the cells’ resistance to very high temperature. The magnitude of this effect depended on the extent and duration of the inhibition of proteolysis. A similar induction of hsps and thermotolerance was seen with another proteasome inhibitor, clasto-lactacystin β-lactone, but not with an inhibitor of vacuolar proteases. Surprisingly, when the reversible inhibitor MG132 was removed, thermotolerance decreased rapidly, while synthesis of hsps continued to increase. In addition, exposure to MG132 and 37°C together had synergistic effects in promoting thermotolerance but did not increase hsp expression beyond that seen with either stimulus alone. Although thermotolerance did not correlate with hsp content, another thermoprotectant trehalose accumulated upon exposure of cells to MG132, and the cellular content of this disaccharide, unlike that of hsps, quickly decreased upon removal of MG132. Also, MG132 and 37°C had additive effects in causing trehalose accumulation. Thus, the resistance to heat induced by proteasome inhibitors is not just due to induction of hsps but also requires a short-lived metabolite, probably trehalose, which accumulates when proteolysis is reduced.  相似文献   

13.
The exposure of exponentially growing BHK cells to supranormal temperatures (41–44 °C, for 15 min to 1 h) induces the synthesis of a new set of proteins, the heat shock proteins, while the synthesis of proteins made before heat shock is repressed at 43 °C. Among the two major heat shock proteins induced, of molecular weight 70 K and 68 K, only the 70 kD protein is found bound to the nuclear matrix. This protein is resolved differently from the normal matrix proteins by isoelectric focusing and, when blotted, does not react with antibodies directed against nuclear matrices. These results show that the 70 kD heat shock protein is a new protein transferred from the cytoplasm to the nucleus, where it binds to the nuclear matrix, suggesting a structural role for this protein.  相似文献   

14.
Multilamellar liposomes prepared from total lipids of red blood cells are agglutinable by the addition of soybean lectin. At 5 °C the rate of agglutination is significantly slower than at 37 °C, in contrast to erythrocyte ghosts and ghosts sonicated to 1 μ vesicles. The slower lateral mobilities of the lectin glycolipid receptor in the lipid liposomes due to increased microviscosity of the bilayer at the lower temperature, might be one explanation of our agglutination results. However, the opposite temperature dependence seen with ghosts argues for a possible protein modulation of the agglutination reaction.  相似文献   

15.
Wu MT  Wallner SJ 《Plant physiology》1984,75(3):778-780
Using cultured pear (Pyrus communis cv Bartlett) cells, heat tolerance induced by heat shock was compared to that developed during growth at high temperature. After growth at 22°C, cells exposed to 38°C for 20 minutes (heat shock) showed maximum increased tolerance within 6 hours. Cells grown at 30°C developed maximum heat tolerance after 5 to 6 days; this maximum was well below that induced by heat shock. Heat shock-induced tolerance was fully retained at 22°C for 2 days and was only partly lost after 4 days. However, pear cells acclimated at 30°C lost all acquired heat tolerance 1 to 2 days after transfer to 22°C. In addition, cells which had been heat-acclimated by growth at 30°C showed an additional increase in heat tolerance in response to 39°C heat shock. The most striking difference between heat shock and high growth temperature effects on heat tolerance was revealed when tolerance was determined using viability tests based on different cell functions. Growth at 30°C produced a general hardening, i.e. increased heat tolerance was observed with all three viability tests. In contrast, significantly increased tolerance of heat-shocked cells was observed only with the culture regrowth test. The two types of treatment evoke different mechanisms of heat acclimation.  相似文献   

16.
The formation of more than trace amounts of ice in cells is lethal. The two contrasting routes to avoiding it are slow equilibrium freezing and vitrification. The cryopreservation of mammalian oocytes by either method continues to be difficult, but there seems a slowly emerging consensus that vitrification procedures are somewhat better for mouse and human oocytes. The approach in these latter procedures is to load cells with high concentrations of glass-inducing solutes and cool them at rates high enough to induce the glassy state. Several devices have been developed to achieve very high cooling rates. Our study has been concerned with the relative influences of warming rate and cooling rate on the survival of mouse oocytes subjected to a vitrification procedure. Oocytes suspended in an ethylene glycol–acetamide–Ficoll–sucrose solution were cooled to −196 °C at rates ranging from 37 to 1827 °C/min between 20 and −120 °C, and for each cooling rate, warmed at rates ranging from 139 to 2950 °C/min between −70 and −35 °C. The results are unambiguous. If the samples were warmed at the highest rate, survivals were >80% over cooling rates of 187–1827 °C/min. If the samples were warmed at the lowest rate, survivals were near 0% regardless of the cooling rate. We interpret the lethality of slow warming to be a consequence of it allowing time for the growth of small intracellular ice crystals by recrystallization.  相似文献   

17.
Electron transport, using succinate as a substrate, was measured polarographically in mitochondria isolated from Phaseolus vulgaris and P. acutifolius plants at 25°C and 32°C. Mitochondria isolated from P. vulgaris plants grown at 32°C had reduced electron transport and were substantially uncoupled. Growth at 32°C had no effect on electron transport or oxidative phosphorylation in P. acutifolius compared to 25°C grown plants. Mitochondria isolated from 25°C grown P. vulgaris plants measured at 42°C were completely uncoupled. Similarly treated P. acutifolius mitochondria remained coupled. The uncoupling of P. vulgaris was due to increased proton permeability of inner mitochondrial membrane. The alternative pathway was more sensitive to heat than the regular cytochrome pathway. At 42°C, no alternative pathway activity was detected. The substantially greater heat tolerance of P. acutifollus compared to P. vulgaris mitochondrial electron transport suggests that mitochondrial sensitivity to elevated temperatures is a major limitation to growth of P. vulgaris at high temperatures and is an important characteristic conveying tolerance in P. acutifolius.  相似文献   

18.
Acholeplasma laidlawii cells were grown in cholesterol-enriched medium and exposed continuously to either air (control), 4.0 vol.% halothane in air at 1 atm pressure (4% atm halothane), or 80% cyclopropane in oxygen for 24 h at 37°C. Cells grown in the presence of 4% atm halothane or 80% cyclopropane had approximately twice as much membrane cholesterol content/mg protein as the control cells. Cells grown in an anesthetic environment also tended to have a higher membrane cholesterol/phospholipid molar ratio compared to control cells. Membranes isolated from halothane-exposed cells grown in a cholesterol-enriched medium were more ordered at 37°C (measurements were made with no anesthetic present) than membranes from control cells grown in an identically enriched medium. This difference in membrane physical state between control and anesthetic-exposed cells decreased as the temperature decreased, and disappeared at approx. 23°C. Continuous exposure of A. laidlawii to 4% atm halothane or 80% cyclopropane for 24 h did not markedly affect membrane fatty acid composition, either in cells grown on an unsupplemented medium or in cells grown in a medium enriched in myristic, palmitic or stearic acids. These results further support the hypothesis that an increased membrane cholesterol content may play a role in the tolerance or dependence that develops after chronic exposure to anesthetic agents.  相似文献   

19.
Cytological preparations were incubated in 0.07 N NaOH at room temperature or 90% formamide (final salt concentration 2 × SSC) at either 65 °C or 37 °C for 2.5 h to denature guinea pig chromosomes. Chromosomes treated with NaOH or formamide at 65 °C showed a large amount of DNA loss, while chromosomes treated with formamide at 37 °C showed little or no DNA loss. Repeated sequences were isolated from guinea pig DNA and [3H]cRNA was transcribed with Escherichia coli RNA polymerase for in situ hybridization. Localization of the [3H]cRNA occurred in the centromeric regions and C-band positive short arms of almost all of the chromosomes in the NaOH preparations. Chromosomes treated with formamide at 65 °C showed the same grain distribution with a decrease in the number of grains/cluster. Slides incubated in formamide at 37 °C showed localization in only a few chromosomes and the number of grains/cluster was greatly diminished. Thermal denaturation of isolated chromatin indicated that incubation of chromosomes in formamide at 37 °C did not fully denature the DNA. C-bands could be induced by treating slides in formamide at either 65 °C or 37 °C when followed by a “reassociation” in 2 × SSC at 65 °C for 16 h. If the “reassociation” step was omitted, C-bands were found in the 65 °C formamide slides but not the 37 °C formamide slides.  相似文献   

20.
Experiment was conducted with the aim of determining the effect of varying water temperature and ration size on growth and body composition of fry of the common carp, Cyprinus carpio. Common carp fry with an initial body weight (BW) of 0.86 g were fed a diet (34.9% protein, 18.3 KJ/g diet) at four ration sizes 4%, 5%, 6% and 7% of their body weight per day and reared at two water temperatures 28 and 32 °C for 60 days. Fry fed with 6% ration showed the highest mean final body weight at 28 °C. Final body weight was significantly (P<0.05) affected by ration and temperature. Cyprinus carpio fry raised at 28 °C had higher feed efficiency (FE) (44.36%) than the fry reared at 32 °C (40.98%) with 4% ration. Further, feed efficiency decreased with increase in ration levels in both temperatures. Protein efficiency ratio (PER) was higher (1.26) at 28 °C than at 32 °C (1.17). At 6% ration, common carp fry showed highest specific growth rate (SGR) (3.82%/day) at 28 °C as compared with at 32 °C (3.57%/day). A linear increase in protein and lipid contents was evident with increasing ration levels up to 6% body weight at both temperatures 28 and 32 °C. Second-order polynomial regression analysis of weight gain and SGR indicated the breakpoints at ration level 6.04% and 6.08% body weight per day at 28 and 32 °C. Hepatosomatic index (HSI) not affected by temperature and ration size while, viscerosomatic index (VSI) influenced (P<0.05) by ration size and temperature. Based on the above results, it may be concluded that 6% BW/day ration is optimal for growth of Cyprinus carpio fry at both the temperatures 28 and 32 °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