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.
Heterogeneity in radiosensitization by heat was studied using one uncloned and five cloned cell lines isolated from a single tumour of a human melanoma xenograft. Cells from passages 7-12 in vitro were given heat treatments of 42.5 degrees C (45 min), 43.5 degrees C (45 min) or 44.5 degrees C (45 min) immediately after exposure to graded doses of radiation. The survival curves after irradiation alone had similar D0 values but differed in the size of the shoulder. The heterogeneity in heat radiosensitization was reflected in differences in decrease of the D0 values. The thermal enhancement ratios, calculated from the D0 values, were in the ranges 1.2 +/- 0.2-1.7 +/- 0.2 (42.5 degrees C), 1.4 +/- 0.3-2.4 +/- 0.4 (43.5 degrees C) and 2.3 +/- 0.4-3.4 +/- 0.4 (44.5 degrees C). Moreover, at 43.5 degrees C the heterogeneity was also reflected in different modifications of the shape of the survival curves. Two lines showed survival curves with a significant shoulder and a relatively low D0 value whereas two other lines had lost the shoulder almost completely but showed relatively high D0 values. All lines showed survival curves with a broad shoulder after heating at 42.5 degrees C, whereas none of the lines showed survival curves with a significant shoulder after heating at 44.5 degrees C.  相似文献   

2.
The membrane fluidity of cells of human (AG1522 human foreskin fibroblasts), rodent [Chinese hamster ovary (CHO) and radiation-induced mouse fibrosarcoma], and feline (Crandall feline kidney) cell lines after heating at 45 degrees C was measured by flow cytometry. In addition, a heat-resistant variant of radiation-induced mouse fibrosarcoma cells and two heat-sensitive CHO strains were studied. Fluorescence polarization of the plasma membrane probe trimethylammonium-diphenylhexatriene was used as a measure of membrane fluidity. The sensitivity of all cell lines to 45 degrees C hyperthermia was compared. The baseline membrane fluidity varied among the cell lines, but did not correlate with sensitivity to hyperthermia. However, CHO cells, especially the heat-sensitive mutants, had the largest increase in membrane fluidity after heating at 45 degrees C, while the heat-resistant mouse fibrosarcoma variants and Crandall feline kidney cells resisted changes in fluidity. In general, the more resistant the cell line was to killing by heat, the more resistant it was to changes in membrane fluidity.  相似文献   

3.
The cell cycle responses of two exponentially growing subpopulations of cells (clones A and D), originally obtained from a human colon adenocarcinoma to X-irradiation, were studied using centrifugal elutriation. Cell suspensions were separated by changing counter-current flow rate while keeping the rotor speed constant (1600 rpm) and the composition of eluted fractions was determined using flow cytometry. The X-ray sensitivity of unseparated clone D cells was somewhat greater than that of clone A cells (e.g. 10% greater at the 10% level of survival). This difference appeared to be due to a greater value of the alpha parameter (one-hit cell killing), using the linear-quadratic equation in which the relative survival S/S0 = exp - (alpha D + beta D2) with dose (D) in Gy. This finding was confirmed in the cell cycle studies where the alpha parameter was always greater for the clone D cells than for the clone A cells. The beta parameter was essentially the same for both cell lines through the cell cycle.  相似文献   

4.
Abstract The cell cycle responses of two exponentially growing subpopulations of cells (clones A and D), originally obtained from a human colon adenocarcinoma to X-irradiation, were studied using centrifugal elutriation. Cell suspensions were separated by changing counter-current flow rate while keeping the rotor speed constant (1600 rpm) and the composition of eluted fractions was determined using flow cytometry. The X-ray sensitivity of unseparated clone D cells was somewhat greater than that of clone A cells (e.g. 10% greater at the 10% level of survival). This difference appeared to be due to a greater value of the α parameter (one-hit cell killing), using the linear-quadratic equation in which the relative survival S / S 0 = exp – (αD +βD2) with dose (D) in Gy. This finding was confirmed in the cell cycle studies where the α parameter was always greater for the clone D cells than for the clone A cells. The β parameter was essentially the same for both cell lines through the cell cycle.  相似文献   

5.
Previous studies suggested that a 26 kDa protein might play an important role in protein synthesis-independent thermotolerance development in CHO cells. To determine if this phenomenon was universal, four mammalian cell lines, viz., CHO, HA-1, murine Swiss 3T3, and human HeLa, were studied. Cells were heated at 42 degrees C, and the level of 26 kDa protein in the nucleus was measured, together with clonogenic survival and protein synthesis. The results demonstrated that 1) the 26-kDa protein was present in the four different cell lines, and 2) the level of the 26 kDa protein in their nuclei was decreased by 30-70% after heating at 42 degrees C for 1 hr. However, restoration of this protein occurred along with development of chronic thermotolerance. The protein synthesis inhibitor cycloheximide (10 micrograms/ml) neither inhibited the development of chronic thermotolerance nor affected the restoration of the 26 kDa protein in the nucleus. In fact, this drug protected cells from hyperthermic killing and heat-induced reduction of 26 kDa protein in the nucleus. Heat sensitizers, quercetin (0.1 mM), 3,3'-dipentyloxacarbocyanine iodide (DiOC5[3]: 5 micrograms/ml), and stepdown heating (45 degrees C-10 min----42 degrees C), potentiated hyperthermic killing and inhibited or delayed the restoration of the 26 kDa protein to the nucleus. These results support a correlated, perhaps causal relationship between the restoration of the 26 kDa protein and chronic thermotolerance development in four different mammalian cell lines.  相似文献   

6.
The survival rate of V79 Chinese hamster lung fibroblasts treated with heat and freezing was investigated. The hyperthermic behavior of these cells was similar to other reports; Thermotolerance was observed below 43 degrees C during continuous heating. There was no difference in survival rate whether the cells were heated while attached to tissue culture flasks or in suspension, although the pH of the medium during the latter situation could be around 8.2 during hyperthermic treatment. If the cells were frozen and thawed before heating, the pattern of survival did not change significantly from that of heating alone, and thermotolerance was still observed below 43 degrees C. The Arrhinius plot of heat sensitivities between 41 and 45 degrees C demonstrated a straight line, giving an activation energy of 171 kcal/mole. A break on the plot around 42.5 degrees C could not be located, probably due to the lack of experimental points. If the cells were frozen and thawed immediately after heating, the survival rate was lower than that expected from considering hyperthermia and freezing as two independent processes. Possible explanations are provided in the result and discussion section.  相似文献   

7.
Hydrogen-peroxide-resistant Chinese hamster fibroblasts, derived from the HA-1 cell line, were isolated following continuous culturing in the presence of progressively increasing concentrations of hydrogen peroxide. The hydrogen-peroxide-resistant phenotype has been stable for over 360 days following removal from H2O2 stress. These H2O2-resistant cell lines demonstrate increased resistance to hyperthermic cell killing mediated by continuous heating at 43 degrees C but not 45 degrees C. The relationship between mammalian cellular adaptation to oxidative stress mediated by H2O2 and resistance to 43 degrees C hyperthermia is discussed.  相似文献   

8.
To quantitatively relate heat killing and heat radiosensitization, asynchronous or G1 Chinese hamster ovary (CHO) cells at pH 7.1 or 6.75 were heated and/or X-irradiated 10 min later. Since no progression of G1 cells into S phase occurred during the heat and radiation treatments, cell cycle artifacts were minimized. However, results obtained for asynchronous and G1 cells were similar. Hyperthermic radiosensitization was expressed as the thermal enhancement factor (TEF), defined as the ratio of the D0 of the radiation survival curve to that of the D0 of the radiation survival curve for heat plus radiation. The TEF increased continuously with increased heat killing at 45.5 degrees C, and for a given amount of heat killing, the amount of heat radiosensitization was the same for both pH's. When cells were heated chronically at 42.4 degrees C at pH 7.4, the TEF increased initially to 2.0-2.5 and then returned to near 1.0 during continued heating as thermal tolerance developed for both heat killing and heat radiosensitization. However, the shoulder (Dq) of the radiation survival curve for heat plus radiation did not manifest thermal tolerance; i.e., it decreased continuously with increased heat killing, independent of temperature, pH, or the development of thermotolerance. These results suggest that heat killing and heat radiosensitization have a target(s) in common (TEF results), along with either a different target(s) or a difference in the manifestation of heat damage (Dq results). For clinical considerations, the interaction between heat and radiation was expressed as (1) the thermal enhancement ratio (TER), which is the dose of X rays alone divided by the dose of X rays combined with heat to obtain an isosurvival, e.g., 10(-4), and (2) the thermal gain factor (TGF), the ratio of the TER at pH 6.75 to the TER at pH 7.4. Since low pH reduced the rate of development of thermal tolerance during heating at low temperatures, low pH enhanced heat killing more at 42-42.5 degrees C than at 45.5 degrees C where thermal tolerance did not develop. Therefore, the increase in the TGF after chronic heating at 42-42.5 degrees C was greater than after acute heating at 45.5 degrees C, due primarily to the increase in heat killing causing an even greater increase in heat radiosensitization. These findings agree with animal experiments suggesting that in the clinic, a therapeutic gain for tumor cells at low pH may be greater for temperatures of 42-42.5 degrees C than of 45.5 degrees C.  相似文献   

9.
T cell subpopulations were obtained from F12.5 and B245/270D T cell lines during long-term culture. Two altered F12.5 subpopulations proliferated more intensively than the original clone. These two subpopulations of F12.5 constantly expressed CD25 (interleukin 2 receptor α chain) at a high level and exogenously added interleukin 2 (IL-2) enhanced cell death for one of these subpopulations. However, the original clone expressed CD25 only after activation and IL-2 inhibited cell death of the original clone. On the other hand, the altered B245/270D subpopulation lost the antigen-specific proliferation ability. This altered cell line under the stimulation culture did not express CD25 even after activation, although the original line expressed CD25. However, the expression pattern of CD25on the altered cell line at resting state was induced similar to that of the original one. These results suggest that an expression pattern of CD25 can be changed during long-term cultures, accompanied with alteration in response to proliferation and cell death. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

10.
Interaction of hyperthermia and pentamidine in HeLa S-3 cells   总被引:1,自引:0,他引:1  
Pentamidine is similar to rhodamine-123 in chemical structure and state of electron charge, and rhodamine-123 was previously shown to be a hyperthermic sensitizer under appropriate cell culture conditions. The present experiments were carried out to determine whether pentamidine would potentiate hyperthermic cell killing and, if so, under what cultural conditions. Exposure of HeLa cells to pentamidine (80 microM) up to 4 h was not cytotoxic in culture medium in the presence or absence of glucose at 37 degrees C and pH 7.4. Cells in the glucose-deprived medium became progressively sensitive to killing as temperature and drug concentration were increased. On the other hand, there was a moderate level of enhanced cell killing in the glucose-fed medium at 42 degrees C. The enhanced effects of heat by the drug were most pronounced under alkaline pH of the culture medium. The cell kill under acidic pH was far less than that observed at neutral or alkaline pH; these effects may be a result of decreased cellular uptake of pentamidine. Together with our previous results on rhodamine-123 and glycolytic inhibitors, the present data with pentamidine are consistent with the concept that deprivation of cellular energy increases sensitivity to cell killing by heat.  相似文献   

11.
Analysis of published data reveals that the introduction of more basic amino acid residues in the flexible N-terminal region of the human tumour necrosis factor alpha (TNF) molecule indicates a weak but consistent trend towards increased in vitro cytotoxicity, especially when the effect of N-terminal length is taken into account. In our laboratory, a series of TNF analogues with a charge modification in the tip region of the molecule was prepared, and cytotoxicity measured. Similar trends in cytotoxicity with increasing basicity of the TNF analogue were found in this study for two mouse cell lines, L929 and WEHI-164 clone 13-1, as well as for the human line KYM-1D4. For the series of analogues as a whole, a general increase in in vitro cytotoxicity with increasing pI values was not apparent, but some analogues with charge reversal in the tip region, for example, the LK-805 analogue (E107K), exhibited significantly increased cytotoxicity in comparison to native TNF in a range of cell lines, including L929, KYM-1D4-K, WEHI-164 clone 13-1, HEPA 1-6 and EAhy926 cell lines. Experiments with heparinase-pre-treated cells demonstrated that the increased in vitro cytotoxicity of LK-805 is most probably due to interactions with cell surface heparan sulphates that effectively concentrate it before binding to TNF receptors occurs. Examination of structural models of TNF bound to soluble TNF receptor 1 (TNFR1) indicates that simple mutations in the tip region most probably cannot interact with receptor binding sites, and therefore do not directly modulate cytotoxicity.  相似文献   

12.
The effect of amiloride on hyperthermic cell killing of normal and thermotolerant mouse fibroblast LM cells was investigated under normal (pH 7.4) and acidic (pH 6.8) conditions. Amiloride is known to inhibit the Na+/H+ exchanger in the plasma membrane, the main pH regulating mechanism in mammalian cells. The effects of low pH and amiloride on the mouse fibroblasts were qualitatively similar. For normal cells, mainly a reduction of the shoulder of the survival curve was observed, while an increase of the slope of the exponential part of the survival curve was found in thermotolerant cells. When a combination of 3 mmol dm-3 amiloride and low pH was used the effect on the hyperthermic sensitivity of normal and thermotolerant cells was not additive. This may be explained by a similarity in the mechanism of action of the two treatments, viz. inhibition of the Na+/H+ exchange, which is probably complete when 3 mmol dm-3 amiloride is used. The amiloride sensitivity of normal and thermotolerant fibroblasts is dose dependent in the range of 0.1 to 3 mmol dm-3. Because the D0 of control cells is almost independent of the amount of amiloride, a concentration-dependent reduction of the thermotolerance ratio is found, especially at higher concentrations of amiloride.  相似文献   

13.
Previous studies with cultured human normal fibroblasts indicated that pretreatment of the cells with zinc for 12 h prior to exposure to the alkylating agent melphalan increased survival by seven- to ninefold over survival values obtained in cultures treated with drug only. Comparable pretreatment of cells derived from a variety of human tumors resulted in an increase in survival of 1.7-fold or less. To determine whether the limited responsiveness to zinc represented a general property of tumor cells (which would be characterized by a lack of highly zinc-responsive subpopulations contained within the parental tumor populations), a series of clones was prepared from the A101D human melanoma line and the A549 human alveolar cell carcinoma line. Cells from each clone were then challenged with melphalan with and without zinc pretreatment. Twenty-five percent of the tumor clones exhibited increased resistance to melphalan following pretreatment with zinc (range of 2.1- to 5.2-fold increase in survival), indicating that the parental tumor lines were highly heterogenous in regard to inducibility to a state of reduced sensitivity to melphalan. There was no evidence of a relationship between zinc-induced reductions in toxicity and induced elevations in total intracellular glutathione content, indicating that the primary effect of zinc is not directed toward elevating intracellular levels of glutathione.  相似文献   

14.
Time-temperature analyses of durations of heating required to achieve isosurvival were used to compare hyperthermic cell killing of synchronous Chinese hamster ovary (CHO) cells heated in G1 or S at temperatures of 42 to 45.5 degrees C. G1 populations were obtained by incubation of mitotic cells for 90 min at 37 degrees C. S phase populations were obtained by incubation of mitotic cells for 12 h at 37 degrees C in medium supplemented with 2 micrograms/ml aphidicolin, a reversible inhibitor of DNA alpha polymerase; S phase survival was also determined in an aphidicolin-free system by using high specific activity [3H]thymidine. In both systems, the thermosensitivity was similar and decreased as the cells progressed from early S phase, in agreement with earlier studies (R. A. Read, M. H. Fox, and J. S. Bedford. Radiat. Res. 98, 491-505 (1984]. A comparison of Arrhenius plots of the inverse of durations of heating required to achieve isosurvival for cells heated in G1 or S phase showed similar temperature dependence above 43.5 degrees C, yet the plots for heat-sensitive S phase cells were offset from those for heat-resistant G1 cells by about 1.5 degrees C, i.e., S phase cells respond to 43 degrees C with a rate similar to that observed in G1 cells heated at 44.5 degrees C. Using least-squares regression of the semilog plots, the curves were analyzed either as continually bending curves or as two straight lines with a break at 43.5 degrees C. When the data were analyzed using two straight lines, no significant differences in the slopes of the time-temperature plots of G1 or S phase cells were observed. A quantitative comparison between the two methods of data analysis demonstrated that in both phases the data were better fit with a continuously curving line, rather than two straight lines.  相似文献   

15.
The interaction between hyperthermia and X irradiation was determined in cultured Reuber H35 hepatoma cells with different states of thermosensitivity. Incubation at 41 degrees C followed by 4-Gy X rays resulted after 2 hr in a stabilization of cell survival for heat or plus X rays, with a maximum synergism factor of 1.6. Thermotolerance did not develop during incubation at 41.7 or 42.5 degrees C. When heat treatment of cells was followed by irradiation, the synergism factor for thermal radiosensitization increased with both the amount of thermal cell killing and the amount of X-ray cell killing; the influence of thermal exposure on the synergism factor was greater than that of the X-ray dose. Cells were made thermotolerant either by incubation at 42.5 degrees C for 30 or 60 min followed by an interval at 37 degrees C, or by continuous incubation at 41 degrees C. In both cases thermotolerance was measured by incubation at 42.5 degrees C. No difference was observed between the maximum thermotolerance achieved with both methods. When cells were irradiated in addition to the second heat treatment, thermal radiosensitization was strongly reduced concomitant with the decreased sensitivity to killing by heat.  相似文献   

16.
The effects of 45 degrees C hyperthermia and gamma radiation have been studied in three normal human fibroblast lines (GM38, GM730, WI38) and compared to the effects in two lines derived from patients with the hereditary disease ataxia telangiectasia (AT3BI, AT5BI). All lines, both normal and gamma-sensitive AT, showed a similar resistance to killing by heat alone, suggesting that the defect responsible for the increased radiation sensitivity in AT lines does not confer increased heat sensitivity. Shouldered survival curves were obtained in each case indicating the ability to accumulate sublethal heat damage. All normal and AT cell lines exhibited increased resistance to the lethal effects of heat in response to a thermal stress, indicating that the defect that causes radiosensitivity in AT cell lines does not prevent the induction of thermotolerance. Heat (45 degrees C, 30 min) was shown to increase the sensitivity of the normal cell lines to killing by gamma radiation. The thermal enhancement ratios obtained ranged from about 2.5 to 3.0. The same heat treatment, however, produced very little increase in the radiation sensitivity of the AT cells. Thermal enhancement ratios of about 1.2 were obtained in these lines. We hypothesize that, in normal cells, this heat treatment inactivates the process which is already defective in AT lines, and that this process may be required for the proper rejoining of double-strand breaks produced during the repair of other radiation-induced lesions.  相似文献   

17.
Possible relations between hyperthermic inactivation of alpha and beta DNA polymerase activity and hyperthermic cell killing or hyperthermic radiosensitization were investigated. Ehrlich Ascites Tumor (EAT) cells and HeLa S3 cells were treated with fractionated doses of hyperthermia. The heating schedules were chosen such that the initial heat treatment resulted in either thermotolerance or thermosensitization (step-down heating) for the second heat treatment. The results show that for DNA polymerase activity and heat radiosensitization (cell survival) no thermotolerance or thermosensitization is observed. Thus hyperthermic cell killing and DNA polymerase activity are not correlated. The correlation of hyperthermic radiosensitization and DNA polymerase activity was substantially less than observed in previous experiments with normotolerant and thermotolerant HeLa S3 cells. We conclude that alpha and beta DNA polymerase inactivation is not always the critical cellular process responsible for hyperthermic cell killing or hyperthermic radiosensitization. Other possible cellular systems that might determine these processes are discussed.  相似文献   

18.
Summary Lymphocytes infiltrating human ovarian carcinoma obtained directly from the tumour mass (tumour-infiltrating lymphocytes, TIL) or from the carcinomatous ascites (tumour-associated lymphocytes, TAL) were expanded in vitro in long-term cultures with interleukin-2 and tested for their specific cytolytic activity. Killing of the autologous tumour was detected only in a proportion of the patients, less frequently in TIL compared to TAL. In fact two out of ten TIL and four out of nine TAL cultures tested showed significant levels of lysis against the autologous tumour. This cytotoxic activity was not restricted to the autologous tumour, as other tumour cell lines, including non-ovarian ones, were lysed as well. The cultures that were not cytotoxic against the autologous tumour were in most cases able to lyse other tumour cell lines of ovarian or other histology. Cloning of TIL from one patient was performed: of 22 clones tested, 4 displayed higher cytotoxicity against the autologous tumour compared to the uncloned population and 3 out of these 4 did not kill an irrelevant carcinoma cell line. In order to stimulate the expansion of putative specific effectors we performed mixed lymphocyte/tumour cultures (MLTC) with autologous or allogeneic tumour cells. No stimulation of cytotoxicity against the autologous tumour was detected after MLTC in nine different TAL populations, using autologous or allogeneic tumours as stimulators. On the contrary, peripheral blood lymphocytes from two patients after MLTC with the autologous tumour showed increased killing of the autologous and decreased killing of an allogeneic target. In conclusion TIL and TAL from ovarian carcinoma expanded in vitro with interleukin-2 usually have non-MHC-restricted cytotoxicity and variable degrees of reactivity against the autologous tumour. A preferential killing for the autologous tumour was not observed even after MLTC. These results do not exclude the existence of tumour-specific cytotoxic T lymphocytes in ovarian carcinoma; nevertheless they suggest that putative specific effectors have very low frequency and that culture techniques for expanding their growth more selectively are still to be optimized.  相似文献   

19.
Four primary antioxidant enzymes were measured in both human and rat glioma cells. Both manganese-containing superoxide dismutase (MnSOD) and copper-zinc-containing superoxide dismutase (CuZnSOD) activities varied greatly among the different glioma cell lines. MnSOD was generally higher in human glioma cells than in rat glioma cells and relatively higher than in other tumor types. High levels of MnSOD in human glioma cells were due to the high levels of expression of MnSOD mRNA and protein. Heterogeneous expression of MnSOD was present in individual glioma cell lines and may be due to subpopulations or cells at different differentiation stages. Less difference in CuZnSOD, catalase, or glutathione peroxide was found between human and rat glioma cells. The human glioma cell lines showed large differences in sensitivity to the glutathione modulating drugs 1,3-bis (2-chloroethyl)-1-nitrosourea (BCNU) and buthionine sulfoximine (BSO). A good correlation was found between sensitivity to BCNU and the activities of catalase in these cell lines. Only one cell line was sensitive to BSO and this line had low CuZnSOD activity.  相似文献   

20.
M-14 human tumor cells have been subjected to two regimens of step-down heating (SDH) consisting of a conditioning treatment at 42 degrees C for 1 h or at 44.5 degrees C for 20 min, immediately followed by heating at 40 degrees C. Both conditioning treatments thermosensitize the cells towards the subsequent heating at 40 degrees C; the thermosensitization ratio is 6.4 for cells conditioned at 42 degrees C for 1 h and 32.3 for cells conditioned at 44.5 degrees C for 20 min. The overall protein synthetic activity is reduced to 32.7% or 18.4% of control values following 1 h at 42 degrees C and 20 min at 44.5 degrees C, respectively; this inhibition is followed by a full recovery of the synthetic activity during the subsequent exposure at 40 degrees C. SDH-treated cells synthetize four heat shock proteins, with approximate molecular weights of 28, 64, 70 and 90 kDa. The pattern of HSPs induction observed in SDH-treated cells is similar to that found in cells subjected to single hyperthermic exposures. Cells subjected to the SDH sequence 42 degrees C/1 h-->40 degrees C/4 h develop thermotolerance, as indicated by a reduced sensitivity to further hyperthermic challenges.  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