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 15 毫秒
1.
To investigate whether sublethal heat shock protects Perkinsus marinus (Dermo)-infected oysters Crassostrea virginica from lethal heat stress, and the effects of P. marinus infection on sublethal heat shock response, oysters were first experimentally challenged with P. marinus. Then, when infections in oysters progressed to moderate levels (parasite burden = 10(4) to 10(5) cells g(-1) wet tissue weight), oysters were treated with a sublethal heat shock at 40 degrees C for 1 h (heat shock + Dermo challenge). Other treatment groups included heat-shocked, unchallenged (non-P. marinus challenged) oysters and non-heat-shocked, P. marinus-challenged and -unchallenged oysters. Thermal tolerance was compared among these treatments by administering a lethal heat treatment at 44 degrees C for 1 h, 7 d after sublethal heat shock. Sublethal heat shock enhanced survival to lethal heat treatment in both P. marinus-challenged and -unchallenged oysters. Although levels of hsp70 isoforms (hsp69 and hsp72) did not vary significantly by heat shock or infection with P. marinus, responses due to these treatments were apparent when comparing hsp70 levels within infected and uninfected oysters. Infection enhanced expression of hsp69, regardless of whether oysters were heat shocked or not. In uninfected oysters, hsp72 increased due to heat shock 2 and 7 d post heat shock. Overall, this study demonstrates that heat shock can improve survival in oysters, even in oysters infected with P. marinus. Expression of hsp70 varied among isoforms after sublethal and lethal heat shocks and in infected and uninfected oysters. The heat shock response was not negatively affected by P. marinus infection.  相似文献   

2.
Mammalian cells transformed with either 9,10-dimethyl-1,2-benzanthracene, SV40 or H-ras oncogene dramatically changed their ability to synthesize DNA and RNA and metabolize polyphosphate when L-glutamine was withdrawn from the growth medium or when heat shocked (growth at 42 degrees C). Untransformed, DNA and RNA synthesis decreased by 50-80% when glutamine was withdrawn, but polyphosphate accumulated whether or not glutamine was supplied. Heat shock did not alter this response. Transformed isogenic cells responded differently; at 37 degrees C, they decreased their synthesis of DNA and RNA if starved for glutamine, whereas at 42 degrees C, synthesis was optimal without glutamine. Transformed cells accumulated polyphosphate at 37 degrees C when starved for glutamine, but at 42 degrees C, no polyphosphate accumulated. This apparent non-dependence on glutamine by transformed cells when heat shocked was found to be due to the production of glutamine from serum proteins through induction of a protease(s).  相似文献   

3.
The occurrence of thermotolerance, induced by an initial heat treatment at 42 degrees C for 30 min, was studied in adult non-proliferating rat hepatocytes in primary culture. Heat treatment at 42 degrees C for 30 min did not affect cell morphology, cell attachment, Na+, K+ pump activity, K+ content and lactate dehydrogenase accumulation into the medium. In contrast, after exposure to 44 degrees C for 30 min a dramatic change in all these parameters was observed. However, of the cells, which remained attached to the substratum 24 h after treatment, Na+, K+ pump activity and K+ content appeared to be normal compared with untreated cells. Cells, pre-treated at 42 degrees C for 30 min, followed by incubation at 37 degrees C for 16 h, were found to be completely thermal resistant against heat treatment at 44 degrees C, as judged by cell morphology, detachment from the substratum, lactate dehydrogenase accumulation, Na+, K+ pump activity and K+ content. These results show that induction and development of thermotolerance can be studied in non-proliferating cells in primary culture.  相似文献   

4.
A strain of Listeria monocytogenes isolated from a drain in a food-processing plant was demonstrated, by determination of D values, to be more resistant to the lethal effect of heat at 56 or 59 degrees C following incubation for 45 min in tryptose phosphate broth (TPB) at pH 12.0 than to that of incubation for the same time in TPB at pH 7.3. Cells survived for at least 6 days when they were suspended in TPB at pHs 9.0, 10.0, and 11.0 and stored at 4 or 21 degrees C. Cells of L. monocytogenes incubated at 37 degrees C for 45 min and then stored for 48 or 144 h in TPB at pH 10.0 were more resistant to heat treatment at 56 degrees C than were cells stored in TPB at pH 7.3. The alkaline-stress response in L. monocytogenes may induce resistance to otherwise lethal thermal-processing conditions. Treatment of cells in 0.05 M potassium phosphate buffer (pH 7.00 +/- 0.05) containing 2.0 or 2.4 mg of free chlorine per liter reduced populations by as much as 1.3 log(10) CFU/ml, while treatment with 6.0 mg of free chlorine per liter reduced populations by as much as 4.02 log(10) CFU/ml. Remaining subpopulations of chlorine-treated cells exhibited some injury, and cells treated with chlorine for 10 min were more sensitive to heating at 56 degrees C than cells treated for 5 min. Contamination of foods by L. monocytogenes cells that have survived exposure to processing environments ineffectively cleaned or sanitized with alkaline detergents or disinfectants may have more severe implications than previously recognized. Alkaline-pH-induced cross-protection of L. monocytogenes against heat has the potential to enhance survival in minimally processed as well as in heat-and-serve foods and in foods on holding tables, in food service facilities, and in the home. Cells surviving exposure to chlorine, in contrast, are more sensitive to heat; thus, the effectiveness of thermal processing in achieving desired log(10)-unit reductions is not compromised in these cells.  相似文献   

5.
Thermotolerance in Escherichia coli is induced by exposing cells to a brief heat shock (42 degrees C for 15 min). This results in resistance to the lethal effect of exposure to a higher temperature (50 degrees C). Mutants defective in the recA, uvrA and xthA genes are more sensitive to heat than the wild-type. However, after development of thermotolerance these mutants are like the wild-type in their heat sensitivity. This suggests that thermotolerance is an inducible response capable of protecting cells from the lethal effects of heat, independently of recA, uvrA and xthA. Thermotolerance does not develop in a dnaK mutant. In addition, the dnaK mutant is sensitive to heat and H2O2, but is resistant to UV irradiation. This implies that the E. coli heat-shock response includes a mechanism that protects cells from heat and H2O2, but not from UV.  相似文献   

6.
The small molecular weight heat shock protein HSP27 was recently shown to confer a stable thermoresistant phenotype when expressed constitutively in mammalian cells after structural gene transfection. These results suggested that HSP27 may also play an important role in the development of thermotolerance, the transient ability to survive otherwise lethal heat exposure after a mild heat shock. In Chinese hamster O23 cells increased thermoresistance is first detected at 2 h after a triggering treatment of 20 min at 44 degrees C, attains a maximum at 5 hours, and decays thereafter with a half-life of 10 h. We found that the development and decay of transient thermotolerance cannot be solely explained on the basis of changes in the cellular concentration of HSP27. The cellular HSP27 concentration is not increased appreciably at 2 h after heat shock and attains a maximum at 14 h. Similar results were obtained in the case of another heat shock protein, HSP70. HSP70 follows slightly faster kinetics of accumulation (peaks at 10 h) and decays much more rapidly (ti/2 = 4h) than HSP27 (t1/2 = 13h). HSP27 has 3 isoelectric variants A, B, and C of which B and C are phosphorylated. In cells maintained at normal temperature, HSP27A represents more than 90% of all HSP27. Shifting the cell culture temperature from 37 to 44 degrees C induces the incorporation of 32P into the more acidic B and C forms, a process that occurs very rapidly since the reduction in the concentration of the A form and a corresponding increase in the level of B and C is detectable by immunoblot analysis within 2.5 min at 44 degrees C. Analyses performed at various times during development and decay of transient thermotolerance revealed a close relationship between the effect of heat shock on HSP27 phosphorylation and cell ability to survive. For example, fully thermotolerant cells (5 h post-induction) are refractory to induction of HSP27 phosphorylation by a 20-min heat shock. The induction of HSP27 phosphorylation was also studied in a family of clonal cell lines of O23 cells that are thermoresistant as a result of the constitutive expression of a transfected human HSP27 gene. In these thermoresistant cells, phosphorylation of the endogenous hamster HSP27 is induced to a level comparable to that found in the thermosensitive parental cells. However, phosphorylation of the exogenous human protein, which represents more than 80% of total HSP27 in these cells, was much less induced.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
When Saccharomyces cerevisiae cells grown at 23 degrees C were transferred to 36 degrees C, they initiated synthesis of heat shock proteins, acquired thermotolerance to a lethal heat treatment given after the temperature shift, and arrested their growth transiently at the G1 phase of the cell division cycle. The bcy1 mutant which resulted in production of cyclic AMP (cAMP)-independent protein kinase did not synthesize the three heat shock proteins hsp72A, hsp72B, and hsp41 after the temperature shift. The bcy1 cells failed to acquire thermotolerance to the lethal heat treatment and were not arrested at the G1 phase after the temperature shift. In contrast, the cyr1-2 mutant, which produced a low level of cAMP, constitutively produced three heat shock proteins and four other proteins without the temperature shift and was resistant to the lethal heat treatment. The results suggest that a decrease in the level of cAMP-dependent protein phosphorylation results in the heat shock response, including elevated synthesis of three heat shock proteins, acquisition of thermotolerance, and transient arrest of the cell cycle.  相似文献   

8.
The kinetic parameters of net exit of D-glucose from human red blood cells have been measured after the cells were loaded to 18 mM, 75 mM and 120 mM at 2 degrees C and 75 mM and 120 mM at 20 degrees C. Reducing the temperature, or raising the loading concentration raises the apparent Km for net exit. Deoxygenation also reduces the Km for D-glucose exit from red blood cells loaded initially to 120 mM at 20 degrees C from 32.9 +/- 2.3 mM (13) with oxygenated blood to 20.5 +/- 1.3 mM (17) (P less than 0.01). Deoxygenation increases the ratio Vmax/Km from 5.29 +/- 0.26 min-1 (13) for oxygenated blood to 7.13 +/- 0.29 min-1 (17) for deoxygenated blood (P less than 0.001). The counterflow of D-glucose from solutions containing 1 mM 14C-labelled D-glucose was measured at 2 degrees C and 20 degrees C. Reduction in temperature, reduced the maximal level to which labelled D-glucose was accumulated and altered the course of equilibration of the specific activity of intracellular D-glucose from a single exponential to a more complex form. Raising the internal concentration from 18 mM to 90 mM at 2 degrees C also alters the course of equilibration of labelled D-glucose within the cell to a complex form. The apparent asymmetry of the transport system may be estimated from the intracellular concentrations of labelled and unlabelled sugar at the turning point of the counterflow transient. The estimates of asymmetry obtained from this approach indicate that there is no significant asymmetry at 20 degrees C and at 2 degrees C asymmetry is between 3 and 6. This is at least 20-fold less than predicted from the kinetic parameter asymmetries for net exit and entry. None of the above results fit a kinetic scheme in which the asymmetry of the transport system is controlled by intrinsic differences in the kinetic parameters at the inner and outer membrane surface. These results are consistent with a model for sugar transport in which movement between sugar within bound and free intracellular compartments can become the rate-limiting step in controlling net movement into, or out of the cell.  相似文献   

9.
10.
Ohtsubo, T., Igawa, H., Saito, T., Matsumoto, H., Park, H. J., Song, C. W., Kano, E. and Saito, H. Enhancement of Cell Killing by Induction of Apoptosis after Treatment with Mild Hyperthermia at 42 degrees C and Cisplatin. Radiat. Res. 156, 103-109 (2001).We examined the interactive effects of cisplatin (1.0 microg/ml) combined with hyperthermia on cell killing and on the induction of apoptosis in IMC-3 human maxillary carcinoma cells. The cytotoxic effects of hyperthermia on IMC-3 cells at 44 degrees C were greater than at 42 degrees C, as has been reported for many other cells. The induction of apoptosis, DNA fragmentation and poly(ADP-ribose) polymerase cleavage were greater after hyperthermia at 44 degrees C for 30 min compared with treatment at 42 degrees C for 105 min, even though both of these heat doses were isoeffective in reducing cell survival to 50%. Treatment with cisplatin at 37 degrees C for up to 120 min did not result in cytotoxicity or the induction of apoptosis. The enhancement ratio for treatment with cisplatin at 42 degrees C was greater than that at 44 degrees C. More apoptosis was induced after the treatment with cisplatin at 42 degrees C compared to treatment with cisplatin at 44 degrees C. Taking these findings together, the combination of cisplatin and hyperthermia at 42 degrees C appeared to be more effective than cisplatin with hyperthermia at 44 degrees C for the induction of apoptosis in IMC-3 cells.  相似文献   

11.
In mouse cells, the major inducible heat shock protein is a protein of 68,000 daltons (hsp68). We have previously shown that mouse plasmacytomas do not express hsp68. We have now made use of these natural mutants to assess the contribution of hsp68 to acquired thermotolerance. An endpoint limiting dilution assay was used to quantify cell survival to lethal stresses. Two test plasmacytoma cell lines (C1.18.1 and J558) and an hsp68-positive myeloma, XC1.1/51, used as a control, were examined. All showed recovery when pretreated for 10 min at 44 degrees C 2 h before exposure to otherwise lethal stresses of 1 to 4 h at 43 degrees C. Similar results were obtained with the Friend erythroleukemia line D1B, which we have also shown not to express hsp68. These results indicate that hsp68 is not required for protection against thermal stresses in mouse cells.  相似文献   

12.
The kinetics of thermal adaptation at the nonlethal temperature of 40 degrees C was studied in CHO (Chinese hamster ovary) cells in vitro. Thermal resistance, demonstrated as an increase in mean 45 degrees C killing time or as an increase in the shoulder of the 45 degrees C survival curve, was fully developed by 2 h. Control cells in early logarithmic phase were more heat sensitive than those in stationary phase. Corresponding 45 degrees C killing time frequency distributions were unimodal with an increase in mean killing time from early logarithmic to stationary phase. Cells which were thermally adapted at 40 degrees C for 6 h had biphasic 45 degrees C killing time frequency distributions, and as cells progressed from early logarithmic to stationary phase the heat-sensitive subpopulation progressively declined. Exposure to 40 degrees C produced a 30% increase in total protein synthesis. Proteins with molecular weights 72, 89, and 109 kDa which correspond to those induced by lethal heat shock were synthesized at 40 degrees C, but there was no close temporal correlation between the development of heat resistance at 40 degrees C and synthesis of the heat shock proteins. Cycloheximide (100 micrograms/ml) reduced the mean 45 degrees C killing time but did not totally prevent the development of heat resistance at 40 degrees C.  相似文献   

13.
The analysis of proteins synthesized in rat thymocytes and mouse teratocarcinoma PCC-4 Aza 1 and myeloma Sp2/0 cells after 1 h of treatment at 42 or 44 degrees C was carried out. Shock at 42 degrees C reduced the total synthetic rate of proteins in all three cell lines and induced "classical" heat-shock protein with a mass of 70 kDa (hsp 70). Heat shock at 44 degrees C resulted in almost complete inhibition of protein synthesis; only a small amount of hsp 70 was synthesized. Meanwhile a new 48-kDa polypeptide (pI = 7.5) was found in the cells exposed to severe heat shock. This protein was compared by peptide mapping with other known polypeptides of the same size: heat-shock protein from chicken embryo cells and mitogen-stimulated polypeptide from human lymphoid cells. The peptide maps were not identical. It was also shown that after a shock at 44 degrees C teratocarcinoma cells were able to accumulate anomalous amounts of hsp 70 despite hsp 70 synthesis inhibition. The data show that reaction of various cells to extreme heat shock depends heavily on cell type.  相似文献   

14.
The metabolism of radioactively labelled D-glucose, L-glutamine, and L-glutamate has been determined in a glycolytic mutant of Chinese-hamster ovary cells, R1.1.7, and in its parent, CHO-K1. The complete oxidation of glucose via the TCA-cycle is negligible in both cell types, but there is significant oxidation of carbon-1. CHO-K1 cells derive most of their energy from glycolysis and are independent of respiration in the short term. R1.1.7 cells are respiration-dependent and are rapidly killed by respiratory inhibitors. Both cell types oxidize L-glutamine and L-glutamate, but the oxidation of these substrates does not appear sufficient to satisfy completely the energy requirements of R1.1.7 cells.  相似文献   

15.
Increased heat tolerance is most often associated with the synthesis of heat-shock proteins following pre-exposure to a nonlethal heat treatment. In this study, a bromegrass (Bromus inermis Leyss cv Manchar) cell suspension cultured in a medium containing 75 microM abscisic acid (ABA) without prior heat treatment had a 87% survival rate, as determined by regrowth analysis, following exposure to 42.5 degrees C for 120 min. In contrast, less than 1% of the control cells survived this heat treatment. The heat tolerance provided by treatment with 75 microM ABA was first evidenced after 4 d of culture and reached a maximum tolerance after 11 d of culture. Preincubation with sucrose partially increased the heat tolerance of control cells and rendered ABA-treated cells tolerant to 45 degrees C for 120 min (a completely lethal heat treatment for control cells). Comparative two-dimensional polyacrylamide gel electrophoresis of cellular protein isolated from heat-tolerant cells identified 43 ABA-responsive proteins of which 26 were heat stable (did not coagulate and remained soluble after 30 min at 90 degrees C). Eight heat-stable, ABA-responsive proteins ranging from 23 to 45 kD had similar N-terminal sequences. The ABA-responsive (43-20 kD), but none of the control heat-stable, proteins cross-reacted to varying degrees with a polyclonal antibody directed against a conserved, lysine-rich dehydrin sequence. A group of 20- to 30-kD heat-stable, ABA-responsive proteins cross-reacted with both the anti-dehydrin antibody and an antibody directed against a cold-responsive winter wheat protein (Wcs 120). In ABA-treated cells, there was a positive correlation between heat- and pH-induced coagulation of a cell-free homogenate and the heat tolerance of these cells. At 50 degrees C, control homogenates coagulated after 8 min, whereas cellular fractions from ABA-treated cells showed only marginal coagulation after 15 min. In protection assays, addition of heat-stable, ABA-responsive polypeptides to control fractions reduced the heat-induced coagulation of cell-free homogenates. Sucrose (8%) alone and control, heat-stable fractions enhanced the thermostability of control fractions, but the most protection was conferred by ABA-responsive, heat-stable proteins in combination with sucrose. These data suggest that stress-tolerance mechanisms may develop as a result of cooperative interactions between stress proteins and cell osmolytes, e.g. sucrose. Hypotheses are discussed implicating the role of these proteins and osmolytes in preventing coagulation and denaturation of cellular proteins and membranes.  相似文献   

16.
The influence of different sporulation temperatures (30, 37, 44 and 52 degrees C) upon heat resistance of Bacillus subtilis was investigated. Heat resistance was greater after higher sporulation temperatures. Relation of heat resistance and temperature of sporulation was not linear over all the range of temperatures tested. Heat resistance increased about tenfold in the range of 30-44 degrees C. Sporulation at 52 degrees C did not show any further increase in heat resistance. This effect was constant over all the range of heating temperatures tested (100-120 degrees C). z value remained constant (z = 9 degrees C). Greater heat resistances at higher temperatures of sporulation were not due to selection of more heat resistant cells by a higher sporulation temperature. Spores obtained from cells incubated at 32 or 52 degrees C always possessed heat resistances that corresponded to the sporulation temperature regardless of the incubation temperature of their vegetative cells.  相似文献   

17.
The relation between rates of protein synthesis and antibody production was studied for hybridoma cells treated at 42 degrees and 44 degrees C. Both the biosynthetic parameters were shown to recover after a mild heat shock at 42 degrees C with approximately the same kinetics. The treatment at 44 degrees C led to a full inhibition of Ig production, and the protein electrophoretic pattern was not recovered to normal state within 4-6 hours. The synthesis of heat shock proteins (HSP) was found only after the treatment at 42 degrees C. It is suggested that the expression of HSP is necessary for the recovery of hybridoma cell activities.  相似文献   

18.
The interaction of heat and X irradiation was studied in normal and polyunsaturated fatty acid (PUFA) substituted mouse fibroblast LM cells. As a result of the substitution the membranes of the PUFA cells were more fluid than the membranes of the normal cells. Three different heat doses were applied (60 min 42 degrees C, 20 min 43 degrees C, and 10 min 44 degrees C) in combination with single or split doses of X rays. Heat radiosensitization was the largest for the 60 min 42 degrees C treatment. Heat radiosensitization and the heat-induced inhibition of the rate of sublethal damage repair were the same for the normal and the PUFA cells. It is concluded from the experiments reported that the processes of hyperthermic inhibition of SLD repair and hyperthermic radiosensitization are independent of membrane fluidity and membrane fatty acid composition.  相似文献   

19.
A flow cytometry (FCM) assay using SYBRgreen and propidium iodide double staining was tested to assess viability and morphological parameters of Perkinsus marinus under different cold- and heat-shock treatments and at different growth phases. P. marinus meront cells, cultivated at 28 degrees C, were incubated in triplicate for 30 min at -80 degrees C, -20 degrees C, 5 degrees C, and 20 degrees C for cold-shock treatments and at 32 degrees C, 36 degrees C, 40 degrees C, 44 degrees C, 48 degrees C, 52 degrees C, and 60 degrees C for heat-shock treatments. A slight and significant decrease in percentage of viable cells (PVC), from 93.6% to 92.7%, was observed at -20 degrees C and the lowest PVC was obtained at -80 degrees C (54.0%). After 30 min of heat shocks at 40 degrees C and 44 degrees C, PVC decreased slightly but significantly compared to cells maintained at 28 degrees C. When cells were heat shocked at 48 degrees C, 52 degrees C, and 60 degrees C heavy mortality occurred and PVC decreased to 33.8%, 8.0%, and 3.4%, respectively. No change in cell complexity and size was noted until cells were heat shocked at >or=44 degrees C. High cell mortality was detected at stationary phase of P. marinus cell culture. Cell viability dropped below 40% in 28-day-old cultures and ranged 11-25% in 38 to 47-day-old cultures. Results suggest that FCM could be a useful tool for determining viability of cultured P. marinus cells.  相似文献   

20.
Four heat-resistant variants were isolated after treatment of Chinese hamster lung cells with the mutagen ethyl methane sulfonate, followed by a single-step selection procedure consisting in a severe hyperthermic treatment of 4 h at 44 degrees C. The isolated clones had a stable resistant phenotype for at least 150 generations during which they showed a 5,000-fold increased survival to a 4-h treatment at 44 degrees C when compared to wild-type cells. Comparative two-dimensional electrophoretic analyses of proteins revealed that, like induced thermotolerant wild-type cells (i.e., cells induced to a transient physiological state of thermotolerance by a sublethal heat conditioning treatment administered 18 h before), the heat-resistant variants had, at normal temperature, an increased content of a heat-shock protein with Mr of 27,000 (HSP27). In three of the four heat-resistant variants, the increased content of HSP27 was correlated with a two-fold increase in the constitutive level of the mRNA encoding HSP27. Chinese hamster HSP27 is composed of three species that differ in their relative isoelectric point, among which the two most acidic forms are phosphoproteins. In both the heat-resistant variant and wild-type cells, heat shock induces a rapid enhancement of the phosphorylation of HSP27: maximal phosphorylation occurs within 10 min upon changing the incubation temperature from 35 degrees to 44 degrees C. A concomitant shift in silver-staining intensity is rapidly detectable between the three isoforms, which seems to indicate that the two phosphorylated species represent post-translational modifications of the more basic species. It is concluded that most likely the enhanced expression of HSP27 is linked to the resistant phenotype of the variants. The study provides supporting evidence that both the content and phosphorylation status of HSP27 are determining factors in the ability of cells to survive hyperthermic treatments.  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