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.
Intracellular particle movements, of both saltatory and streaming types, in HeLa S-3 cells were simultaneously interrupted after 1 h exposure of cells to 43 degrees C, within 10 min at 44 degrees C and within 5 min at 45 degrees C. Intracellular movement inhibited after 15 min at 44 degrees C and 10 min at 45 degrees C was not reversible in cells rescued at 37 degrees C. Brownian motion was not observed in heat-treated cells while they were maintained at elevated temperatures, but became pronounced in blebbing which occurred shortly after they were returned to 37 degrees C. Returning these cells to 45 degrees C intensified the Brownian activity inside blebs, and rapidly induced cell lysis. The same heat-treated cells were simultaneously studied by laser-Doppler microscopy, which confirmed: a) that flow (cytoplasmic streaming) is completely arrested at 44 degrees C within 10 min, b) flow recovered in 10-15 min in cells rescued after 10-15 min at 44 degrees C, c) submicroscopic particles down to the size of water molecules had faster self-diffusion coefficients at 44 degrees C than at 37 degrees C. Proton nmr studies on cells exposed from 4 to 45 degrees C gave corrected relaxation times T1 and T2 which rose with temperature in a predictable manner. Inhibition of cellular movement at elevated temperatures was not specifically attributable to the depletion of intracellular ATP levels.  相似文献   

2.
Shi B  Xia X 《Current microbiology》2003,46(2):120-123
Adaptation to novel environments usually entails morphological changes. The cell morphology of six experimental populations of Pseudomonas pseudoalcaligenes and their common ancestor were examined with scanning electron microscopy (SEM). The six experimental populations were propagated under different temperatures for 10 months: three of them cultured at constant normal temperature (35 degrees C) forming the control group, and the other three cultured at incremental higher temperatures (from 41 degrees to 47 degrees C) as the HT group. SEM showed the deformed and elongated cells in the 6-h cultures of both ancestral and control populations at 45 degrees C, indicating that 45 degrees C is stressful for the ancestral and the control populations. In contrast, the HT populations retained normal cell shape in the 6-h cultures at both 35 degrees C and 45 degrees C. The mean cell volumes of control and HT populations increased 29% and 34%, respectively, relative to the ancestor at their respective thermal regimens, suggestion that the culturing conditions might favor larger cells.  相似文献   

3.
By using fluorescent microscopy and acridine orange staining it was shown in the studies on short-term culture of human cells that the melting patterns of chromatin DNA of intact lymphocytes of healthy individuals represented the curves with 6 maxima (F530) at the temperature ranges of 45 degrees C, 65 degrees C, 85 degrees C, and 92 degrees C (P less than 0.01). The melting patterns of lymphocytes from patients with Down's disease represented curves with 4 maxima at the temperature ranges of 65 degrees C, 85 degrees C, 88 degrees C, 92 degrees C (P less than 0.01). No decline in the fluorescence intensity at the temperature intervals of 78 degrees C-85 degrees C was apparently due to a greater degree of condensation of definite regions of the trisomal cell chromatin complex. Possible mechanisms accounting for structural readjustments of the interphasic human lymphocyte chromatin occurring under thermal effects are discussed.  相似文献   

4.
The temperature-sensitive (ts) Chinese hamster ovary (CHO) cell mutant tsH1 contains a thermolabile leucyl-tRNA synthetase. Upon incubation at the nonpermissive temperature of 39.5 degrees C, the enzyme became reversibly inhibited over a period of minutes, and the cells lost viability over a period of many hours. However, killing of tsH1 by acute heating at 45 degrees C was identical to that of wild-type (SC) cells. In addition, the heat-induced inhibition of protein synthesis was similar for both cell types, as measured after acute heating at 45 degrees C. Furthermore, both killing and inhibition of protein synthesis showed thermotolerance in both cell types. In contrast to the effects at 45 degrees C, at 39.5 degrees C, neither the inhibition of leucyl-tRNA synthetase activity nor the killing of tsH1 expressed thermotolerance. Also, treatment of tsH1 at 39.5 degrees C did not induce thermotolerance to killing at 45 degrees C. The inhibition of leucyl-tRNA synthetase activity in tsH1 at 39.5 degrees C was further distinguished from the 45 degrees C-induced inhibition of protein synthesis in SC cells by a much more rapid reversal of the inhibition of leucyl-tRNA synthetase activity. Also, the rate of reversal of the inhibition of protein synthesis by 45 degrees C in SC cells was decreased by increased heat dose. Such was not true for the 39.5 degrees C inhibition of leucyl-tRNA synthetase activity in tsH1. The data indicate that there exist two distinct types of thermal inhibition--one slowly reversible type which was observed during and after heating at 45 degrees C and both induced and expressed thermotolerance, and a second, rapidly reversible type, which was evident only during heating of tsH1 at 39.5 degrees C and neither induced nor expressed thermotolerance.  相似文献   

5.
We have shown that heat shock does not induce the synthesis of hsp70 in FM3A cells maintained at a low culture temperature of 33 degrees C although it does so in cells maintained at 37 degrees C [T. Hatayama et al. (1991) Biochem. Int. 24, 467-474]. In this paper, we show that FM3A cells maintained at 37 degrees C produced hsp70 mRNA during continuous heating at 42 degrees C or during postincubation at either 37 or 33 degrees C after being heated at 45 degrees C for 15 min, whereas cells maintained at 33 degrees C did not produce hsp70 mRNA during continuous heating at 37, 39, 42, or 45 degrees C, or during postincubation after being heated at any temperature. Thus the lack of hsp70 synthesis in cells maintained at 33 degrees C seemed to be due to the absence of hsp70 mRNA induction. Also, hsp70 was accumulated in cells maintained at 37 degrees C during continuous heating at 42 degrees C and during postincubation at 37 degrees C after heat shock at 45 degrees C, but not during postincubation at 33 degrees C. The cellular level of the constitutive hsp73 as well as the mRNA level were both similar in cells maintained at 33 and 37 degrees C. On the other hand, the cellular level of the constitutive hsp105 in cells maintained at 33 degrees C was only half of that in cells maintained at 37 degrees C. These hsp105 levels increased significantly in both types of cells after continuous heating at 39 degrees C. These findings indicate that the culture temperature affects not only the induction of hsp70 mRNA but also the accumulation of hsp70 and hsp105 in the cells.  相似文献   

6.
Asynchronous Chinese hamster ovary cells were microinjected with glutathione disulfide (GSSG). Successfully injected cells were scored by coinjecting FITC-dextran with GSSG, followed by fluorescent microscopy. After microinjection, cells were incubated for 2.5 h at 37 degrees C to permit thermotolerance development and then heated at 45 degrees C for 40 min. Cellular heat sensitivity was quantitated by counting the number of grains per cell after labeling heated cells with tritiated amino acids and processing for autoradiography. The data show that microinjection of GSSG induced thermotolerance which increased the number of grains per cell up to 500% of controls. Cells that were exposed to similar concentrations of GSSG in culture medium without microinjection or microinjected without GSSG did not develop thermotolerance.  相似文献   

7.
Adherence of yeasts to other microorganisms and epithelial cell surfaces is important in their colonization. Comparative studies based on the coaggregation of Candida dubliniensis versus Candida albicans with Fusobacterium nucleatum and other oral bacteria suggested differences in the surfaces of these yeasts. Transmission electron microscopy was used to test the hypothesis that there are morphologic variations in the cell surface of these two species. C. dubliniensis type strain CD36 and C. albicans ATCC 18804 were grown on Sabouraud's dextrose agar at various growth temperatures. In some experiments suspensions of yeast cells were treated with dithiothreitol. Fixation for transmission electron microscopy was accomplished using dimethylsulfoxide and alcian blue added to 3% paraformaldehyde and 1% glutaraldahyde in cacodylate buffer. The cell wall of both species was predominantly electron lucent and was visibly differentiated into several layers. A thin electron dense outer layer was seen with clearly visible fibrillar structures, closely associated to the cytoplasmic membrane. The length of the fibrils of the C. albicans cells grown at 37 degrees C was approximately two times greater than those of the cells grown at 25 degrees C. The fibrils of the 37 degrees C-grown cells were thin, distinct and tightly packed whereas those of the 25 degrees C-grown cells appeared blunt, loosely spaced and aggregated. C. dubliniensis demonstrated short, blunt fibrils appearing similar to those of the 25 degrees C-grown C. albicans cells. C. dubliniensis showed no difference in the density, length and arrangement of fibrils between the 25 degrees C and 37 degrees C growth temperatures. The shortest and most aggregated fibrils seen were of the 45 degrees C-grown C. albicans cells. Dithiothreitoltreated 37 degrees C-grown C. albicans cells revealed a distorted and partially destroyed fibrillar layer. In this investigation C. dubliniensis, unlike C. albicans, displayed an outer fibrillar layer that did not vary with variations in growth temperature. In addition, the fibrils on the C. dubliniensis cells were similar to those of the 25 degrees C-grown C. albicans in that they were considerably shorter and less dense than those of the 37 degrees C-grown C. albicans cells. It can be postulated, that C. dubliniensis exhibits constant cell surface characteristics consistent with hydrophobicity and that this property may give this species an ecological advantage. Therefore, C. dubliniensis may compete well in oral environments via enhanced attachment to oral microbes and other surfaces, perhaps even more efficiently than C. albicans.  相似文献   

8.
Bacillus subtilis RUB1012 and RUB1013 have the following phenotype when grown at 45 degrees C: no growth on tryptose blood agar base, growth as clumps of spheres in broth culture, a slow autolysis rate, and a low proportion of teichoic acid to peptidoglycan. Revertants of strain RUB1012 (RUB2032, RUB2012, and RUB2042) that could grow on tryptose blood agar base were isolated. Each revertant had a different proportion of teichoic acid to peptidoglycan. The nanomoles of phosphorus per milligram of cell wall at the nonpermissive temperature were 141, 160, 236, and 541 for strain RUB1012 and revertants RUB2032, 2012, and 2042, respectively, as compared with 1,100 for the parent strain. With most bacteriophage tested, plating efficiency was related to the amount of glucosylated teichoic acid. Scanning electron microscopy was used to study strain RUB2032 during a shift from growth at 30 degrees C to growth at 45 degrees C. The change from rod to sphere began with the thickening of the cylindrical portion of the cell. Caps of the cells appeared to be immune to the thickening process. During growth, the cells became progressively shorter and thicker, and cell separation was inhibited. When cells of strain RUB2032 were shifted from growth at 45 degrees C to growth at 30 degrees C, accumulation of an amorphous material on the outer surfaces of the cells preceded the change from sphere to rod morphology. Cells remained clumped, with rods appearing at the periphery of the clumps. Analysis by DNA-mediated transformation and PBS1-mediated transduction indicated that strains RUB1012 and RUB1013 have multiple mutations mapping in the same region as other cell wall mutations.  相似文献   

9.
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   

10.
The effect of exposure to elevated temperatures (41-45 degrees C) on the repair of radiation-induced DNA strand breaks was measured in monolayer cultured Chinese hamster ovary (CHO) cells. Prior exposure of cells to temperatures between 43 and 45 degrees C resulted in significant decreases in the rate of repair of DNA damage. Exposure to 45 degrees C for 15 min slowed the rate of DNA repair to 0.17 of the control repair rate. The To for inactivation of DNA repair was observed to be 34, 13 and 6 min at 43, 44 and 45 degrees C, respectively. Stepdown-heating (45 degrees C for 15 min followed by repair at 41 degrees C) resulted in greater inhibition of DNA repair (0.11 of the control rate) than was observed after acute heating alone. Repair at 41 degrees C was observed to proceed in unheated cells at a faster rate than at 37 degrees C. An Arrhenius analysis of the inactivation kinetics of DNA repair between 43 and 45 degrees C indicated an activation energy of 140 kcal mol-1 of protein for the inhibition of DNA repair. In general, the results were inconsistent with either a retardation of the DNA repair rate or an increase in unrepaired DNA lesions being responsible for heat-induced radiosensitization.  相似文献   

11.
1. Cell-free systems prepared from Krebs II ascites cells incubated at 45 degrees C have a much lower endogenous activity than those from cells incubated at 37 degrees C. The endogenous activity is mainly due to completion of polypeptide chains initiated in the intact cell. The low activity of the 45 degrees C system is due to a lesion in initiation in cells incubated at 45 degrees C. 2. Cell-free systems from cells incubated at 45 degrees C can translate efficiently poly (U) at 8 mM Mg2+. However, they initiate poorly on globin mRNA, indicating that these systems reflect the situation in the intact cell. 3. The lesion in globin mRNA translation in 45 degrees C systems can be overcome by addition of reticulocyte initiation factors. At saturation concentrations of factors, the response of a 45 degrees C system is restored to almost normal. 4. 45 degrees C systems from 40-S initiation complexes with methionyl tRNAfmet almost as efficiently as normal, but their ability for form 80-S complexes with globin mRNA is impaired, unless they are supplied with exogenous initiation factors.  相似文献   

12.
The effect of elevated temperature on transmembrane potential was studied in Chinese hamster ovary cells in vitro using tetraphenylphosphonium cation (TPP+) and 3,3'-dipentyloxacarbocyanine [Di-O-C5(3)], two unrelated lipophilic cation probes that equilibrate across the plasma membrane according to the transmembrane potential. Uptake of TPP+ was measured using a tritium-labeled probe and the uptake of the fluorescent probe Di-O-C5(3) was measured by flow cytometry. The Nernst equation was used to calculate transmembrane potential. The absolute values obtained for transmembrane potential at 37 degrees C using the two probes were different, but qualitatively similar results were obtained using either probe in the hyperthermia studies. Transmembrane potential measured at 43 and 45 degrees C was at least 20% higher than that measured at 37 degrees C, and the difference was statistically significant (P = 0.025 and P less than 0.01, respectively). The hyperpolarization induced by exposure to 45 degrees C persisted temporarily after cells had been returned to 37 degrees C. The hyperpolarization at 37 degrees C associated with a previous exposure to hyperthermia was maximal after cells had been held at 45 degrees C for 2.0 min, and fell to normal levels after 15.0 min at 37 degrees C.  相似文献   

13.
Flow cytometry was used to measure the fluorescence polarization of the lipid probe trimethylammonium-diphenylhexatriene as an indicator of plasma membrane fluidity of Chinese hamster ovary (CHO) cells heated under various conditions. Fluorescence polarization was measured at room temperature about 25 min after heating. When cells were heated for 45 min at temperatures above 42 degrees C, fluorescence polarization decreased progressively, signifying an increase in plasma membrane fluidity. The fluorescence polarization of cells heated at 42 degrees C for up to 55 h was nearly the same as for unheated control populations, despite a reduction in survival. The fluorescence polarization of cells heated at 45 degrees C decreased progressively with heating time, which indicated a progressive increase in membrane fluidity. The fluorescence polarization distributions broadened and skewed toward lower polarization values for long heating times at 45 degrees C. Thermotolerant cells resisted changes in plasma membrane fluidity when challenged with subsequent 45 degrees C exposures. Heated cells were sorted on the basis of their position in the fluorescence polarization distribution and plated to determine survival. The survival of cells which were subjected to various heat treatments and then sorted from high or low tails of the fluorescence polarization histograms was not significantly different. These results show that hyperthermia causes persistent changes in the membrane fluidity of CHO cells but that membrane fluidity is not directly correlated with cell survival.  相似文献   

14.
The major herpes simplex virus DNA-binding protein, ICP8, was purified from cells infected with the herpes simplex virus type 1 temperature-sensitive strain tsHA1. tsHA1 ICP8 bound single-stranded DNA in filter binding assays carried out at room temperature and exhibited nonrandom binding to single-stranded bacteriophage fd DNA circles as determined by electron microscopy. The filter binding assay results and the apparent nucleotide spacing of the DNA complexed with protein were identical, within experimental error, to those observed with wild-type ICP8. Thermal inactivation assays, however, showed that the DNA-binding activity of tsHA1 ICP8 was 50% inactivated at approximately 39 degrees C as compared with 45 degrees C for the wild-type protein. Both wild-type and tsHA1 ICP8 were capable of stimulating viral DNA polymerase activity at permissive temperatures. The stimulatory effect of both proteins was lost at 39 degrees C.  相似文献   

15.
A psychrotrophic pseudomonad isolated from iced fish oxidized alanine at temperatures close to 0 degrees C and grew over the range 0 degrees C-35 degrees C. The rate of oxidation of alanine, measured manometrically, by cells grown at 2 degrees C was lower than that of cells grown at 22 degrees C. However, the consumption of oxygen after heat treatment at 35 degrees for 35 min was reduced considerably by 2 degrees C grown cells. Alanine oxidase activity was tested in an extract from cells grown at 2 degrees C and 22 degrees C with alanine as the sole carbon, nitrogen, and energy source. Cells grown at 2 degrees C produced an alanine oxidase with a temperature optimum of 35 degrees C and pH optimum of 8, which lost about 80% activity by heat treatment at 40 degrees C for 30 min. There was no change in activity after dialysis at pH 7, 8, or 9. Extracts from cells grown at 22 degrees C contained an alanine oxidase system with an optimum temperature of 45 degrees C, a pH optimum above 8, and only about 30% reduction of activity after heat treatment. This enzyme activity was concentrated in the 0.5 M elution fraction from a Sephadex column, and dialysis reduced the activity at pH 7 and 8. Mesophilic enzyme synthesis apparently started around a growth temperature of 10 degrees C. The crude alanine oxidase systems of Pseudomonas aeruginosa derived from cells grown at 13 degrees C and 37 degrees C had a common optimum temperature of 45 degrees C. These data suggest that one mechanism of psychrophilic growth by psychrotrophic bacteria may be the induction of enzymes with low optimum temperatures in response to low temperature conditions.  相似文献   

16.
AIMS: The present work investigates the possibility that temperature could regulate the pattern of transformation of 2,4'-chlorobiphenyl (2,4'-CB) by psychrotolerant Hydrogenophaga sp. IA3-A. Methods AND RESULTS: Transformation of 2,4'-chlorobiphenyl to 2- and 4-chlorobenzoic acid (2- and 4-CBA), and meta-cleavage products by cells of strain IA3-A incubated at 10 degrees C, 25 degrees C, 37 degrees C or 45 degrees C were monitored by UV spectrometry, HPLC and GC-MS analyses. Cultures incubated at 10 degrees C, 25 degrees C or 37 degrees C produced low amounts of CBAs and excess levels of meta-cleavage products from 2,4'-CB. Cultures incubated at 45 degrees C transformed most of the degraded 2,4'-CB to CBAs and low level of meta-cleavage product. Culture extracts contained unusual varieties of isomeric hydroxylated metabolic products. CONCLUSIONS: Efficient transformation of 2,4'-CB to CBAs was possible in cultures incubated at 45 degrees C. Evidence for the involvement of multiple pathways in the transformation of 2,4'-CB in strain IA3-A suggests that differential regulation of the pathways at different temperatures was likely responsible for the change in the pattern of transformation of 2,4'-CB in cultures incubated at 45 degrees C. Significance AND IMPACT OF THE STUDY: It may be possible to condition cells to transform chlorinated biphenyls more efficiently without accumulating excess level of toxic intermediates.  相似文献   

17.
When Chinese hamster ovary (CHO) cells were exposed to 22 degrees C for 2 hr prior to 42.4 degrees C hyperthermia, neither the shoulder region of the survival curve nor the characteristic development of thermotolerance after 3-4 hr of heating were observed. Absolute cell survival after 4 hr at 42.4 degrees C was decreased by a factor of between 10 and 100 (depending on the rate of heating of nonprecooled controls). Conditioning at 30 degrees C for 2 hr, 26 degrees C for 2 hr, or 22 degrees C for 20 min followed by heating to 42.4 degrees C over 30 min did not result in sensitization. Prolonged (16 hr) conditioning at 30 degrees C, however, increased the cytotoxicity of immediate exposure to 41.4 or 45 degrees C with maximum sensitization to 45 degrees C occurring after 6 hr at 30 degrees C. Both 3- and 18-hr pretreatments at 30 degrees C similarly increased the cytotoxicity of 45-41.5 degrees C step-down heating (D0 = 28 min in precooled versus 40 min in nonprecooled cells).  相似文献   

18.
Cultured mouse L cells undergo apoptosis upon 1 h heat shock at 43 and 45 degrees C. Morphologically characteristic apoptotic cells begin to appear soon after the shock. Immunohistochemistry with anti-transglutaminase antibody shows that in most treated cells the enzyme is induced. Its activation results in the formation of highly cross-linked detergent-resistant apoptotic bodies during recovery. Cycloheximide added during hyperthermic stress inhibits the appearance of apoptotic bodies, showing that heat-shock-induced apoptosis is dependent on protein neosynthesis. The analysis of colony-forming ability of heat-shocked L cells shows a survival of 5% at 43 degrees C and less than 0.02% at 45 degrees C. When protein synthesis is inhibited during heat shock the fraction of surviving cells increases to 23% at 43 degrees C and 0.9% at 45 degrees C. This suggest that part of the cells that die upon heat shock are not heavily damaged and would have survived in the presence of a block in protein synthesis.  相似文献   

19.
Thermotolerance and the heat-shock response in Candida albicans   总被引:3,自引:0,他引:3  
At elevated temperatures, yeast cells of Candida albicans synthesized nine heat-shock proteins (HSPs) with apparent molecular masses of 98, 85, 81, 76, 72, 54, 34, 26 and 18 kDa. The optimum temperature for the heat-shock response was 45 degrees C although HSPs were detected throughout the range 41-46 degrees C. Protein synthesis was not observed in cells kept at 48 degrees C. Yeast cells survived exposure to an otherwise lethal temperature of 55 degrees C when they had previously been exposed to 45 degrees C. The thermotolerance induced during incubation at 45 degrees C required protein synthesis, since protection was markedly reduced by trichodermin. Mercury ions induced a set of three stress proteins, one of which corresponded in size to an HSP, and cadmium ions evoked one stress protein seemingly unrelated to the HSPs observed after temperature shift.  相似文献   

20.
When CHO cells were treated either for 10 min at 45-45.5 degrees C or for 1 hr with 100 microM sodium arsenite (ARS) or for 2 hr with 20 micrograms/ml puromycin (PUR-20), they became thermotolerant to a heat treatment at 45-45.5 degrees C administered 4-14 hr later, with thermotolerance ratios at 10(-3) isosurvival of 4-6, 2-3.2, and 1.7, respectively. These treatments caused an increase in synthesis of HSP families (70, 87, and 110 kDa) relative to total protein synthesis. However, for a given amount of thermotolerance, the ARS and PUR-20 treatments induced 4 times more synthesis than the heat treatment. This decreased effectiveness of the ARS treatment may occur because ARS has been reported to stimulate minimal redistribution of HSP-70 to the nucleus and nucleolus. Inhibiting protein synthesis with cycloheximide (CHM, 10 micrograms/ml) or PUR (100 micrograms/ml) after the initial treatments greatly inhibited thermotolerance to 45-45.5 degrees C in all cases. However, for a challenge at 43 degrees C, thermotolerance was inhibited only for the ARS and PUR-20 treatments. CHM did not suppress heat-induced thermotolerance to 43 degrees C, which was the same as heat protection observed when CHM was added before and during heating at 43 degrees C without the initial heat treatment. These differences between the initial treatments and between 43 and 45 degrees C may possibly be explained by reports that show that heat causes more redistribution of HSP-70 to the nucleus and nucleolus than ARS and that redistribution of HSP-70 can occur during heating at 42 degrees C with or without the presence of CHM. Heating cells at 43 degrees C for 5 hr after thermotolerance had developed induced additional thermotolerance, as measured with a challenge at 45 degrees C immediately after heating at 43 degrees C. Compared to the nonthermotolerant cells, thermotolerance ratios were 10 for the ARS treatment and 8.5 for the initial heat treatment. Adding CHM (10 micrograms/ml) or PUR (100 micrograms/ml) to inhibit protein synthesis during heating at 43 degrees C did not greatly reduce this additional thermotolerance. If, however, protein synthesis was inhibited between the initial heat treatment and heating at 43 degrees C, protein synthesis was required during 43 degrees C for the development of additional thermotolerance to 45 degrees C.(ABSTRACT TRUNCATED AT 400 WORDS)  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