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.
Cell spreading on extracellular matrix and associated changes in the actin cytoskeleton (CSK) are necessary for progression through G(1) and entry into S phase of the cell cycle. Pharmacological disruption of CSK integrity inhibits early mitogenic signaling to the extracellular signal-regulated kinase (Erk) subfamily of the mitogen-activated protein kinases (MAPKs) and arrests the cell cycle in G(1). Here we show that this block of G(1) progression is not simply a consequence of inhibition of the MAPK/Erk pathway but instead it reveals the existence of a discrete CSK-sensitive checkpoint. Use of PD98059 to inhibit MAPK/Erk and cytochalasin D (Cyto D) to disrupt the actin CSK at progressive time points in G(1) revealed that the requirement for MAPK/Erk activation lasts only to mid-G(1), while the actin CSK must remain intact up to late G(1) restriction point, R, in order for capillary endothelial cells to enter S phase. Additional analysis using Cyto D pulses defined a narrow time window of 3 h just prior to R in which CSK integrity was shown to be critical for the G(1)/S transition. Cyto D treatment led to down-regulation of cyclin D1 protein and accumulation of the cdk inhibitor, p27(Kip1), independent of cell cycle phase, suggesting that these changes resulted directly from CSK disruption rather than from a general cell cycle block. Together, these data indicate the existence of a distinct time window in late G(1) in which signals elicited by the CSK act independently of early MAPK/Erk signals to drive the cell cycle machinery through the G(1)/S boundary and, hence, promote cell growth.  相似文献   

2.
A H Uggla 《Mutation research》1990,231(2):233-242
Chinese hamster ovary (CHO) cells were synchronized by mitotic shake-off, treated with the fluorochrome acridine orange (AO; 0.5 micrograms/ml), washed free of excess dye and subsequently exposed to visible light (2 X 40 W/8 Wm-2). The light exposure was performed on cells in the G1, G1/S, S or G2 phase of the cell cycle. AO + light induced high frequencies of aberration in the S phase and even higher in the G1 phase. The aberrations observed were all of the chromatid type. The chromosome-type aberrations (dicentrics, rings) obtained when cells in the G1 phase were exposed to X-rays were not found after corresponding treatments with AO + light. With the exception of an increased frequency of gaps, no chromosomal aberrations were induced in G2-phase cells. Sister-chromatid exchanges were efficiently produced by the photodynamic system in the G1, G1/S and S phase of the cell cycle. In other experiments, AO-treated unsynchronized CHO cells were exposed to light in the presence of the hydroxyl radical scavengers mannitol (100 mM) and 5-dimethyl thiourea (100 mM). In parallel experiments these scavengers were found to reduce markedly the chromosome breaking effects by X-rays but had no influence on the photodynamic induction of chromosomal alterations. The results presented show that the visible light-induced chromosomal alterations in CHO cells sensitized with the fluorochrome AO are obtained by an S-dependent mechanism. Furthermore, the results indicate that the hydroxyl free radical does not play a major role in the production of chromosomal alterations by AO + light.  相似文献   

3.
Hydroxyurea treatment affects the G1 phase in next generation CHO cells   总被引:1,自引:0,他引:1  
DNA replication kinetics were studied in populations of synchronized CHO cells treated in the previous generation with hydroxyurea. These CHO cells were re-synchronized by selective detachment of mitotic cells after previously synchronized G1 traversing cultures were treated with 0.1 mM and 2 mM hydroxyurea for 9 and 13 h. Our results show that these cells exhibit a shortening of G1 of at least 1 h relative to cells selected in mitosis from untreated exponentially growing cultures. Survival studies indicated that the hydroxyurea treatments did not affect plating efficiencies. Cell viability was reduced when the initially synchronized populations were blocked with 2 mM, but not 0.1 mM hydroxyurea for greater than 13 h. DNA replication measurements after these blocks showed that all cultures treated with 2 mM hydroxyurea for either 9, 13 or 15 h were blocked at the same point near the G1/S boundary, and then progressed through S phase with similar kinetics. The observed shortening of G1 in the next generation of these cells was independent of both the concentration (0.1 or 2.0 mM) and the time (9 or 13 h) of the hydroxyurea block. These results suggest that specific events relating to the next cell generation can be uncoupled from DNA synthesis and can occur when hydroxyurea inhibits normal cell cycle traverse of G1 cells into and through S phase.  相似文献   

4.
Hydrogen peroxide (H(2)O(2)) induces a number of events, which are also induced by mitogens. Since the progression through the G1 phase of the cell cycle is dependent on mitogen stimulation, we were interested to study the effect of H(2)O(2) on the cell cycle progression. This study demonstrates that H(2)O(2) inhibits DNA synthesis in a dose-dependent manner when given to cells in mitosis or at different points in the G1 phase. Interestingly, mitotic cells treated immediately after synchronization are significantly more sensitive to H(2)O(2) than cells treated in the G1, and this is due to the inhibition of the cell spreading after mitosis by H(2)O(2). H(2)O(2) reversibly inhibits focal adhesion activation and stress fiber formation of mitotic cells, but not those of G1 cells. The phosphorylation of MAPK is also reversibly inhibited in both mitotic and G1 cells. Taken together, H(2)O(2) is probably responsible for the inhibition of the expression of cyclin D1 and cyclin A observed in cells in both phases. In conclusion, H(2)O(2) inhibits cell cycle progression by inhibition of the spreading of mitotic CHO cells. This may play a role in pathological processes in which H(2)O(2) is generated.  相似文献   

5.
R A Coss 《Radiation research》1986,107(1):143-146
The phenomena of heat-induced G1 delay and thermal resistance were compared in synchronous populations of CHO cells. Mildly toxic induction doses of 5 min (Single cell survival, (SCS) = 0.90 +/- 0.06) and 10 min (SCS = 0.69 +/- 0.12) at 45 degrees C resulted in G1 delays of 4.3 and 11.3 h, respectively. Thermal resistance was tested (30 min, 45 degrees C) for up to 32-92 h following the induction dose. Thermal resistance did not start to decay prior to 26 h following the induction dose. These data confirm reports by R. R. Read, M. H. Fox, and J. S. Bedford [Radiat. Res. 98, 491-505 (1984)] and G. L. Rice, J. W. Gray, P. N. Dean, and W. C. Dewey [Cancer Res. 44, 2368-2376 (1984)] that acutely heated G1 populations of CHO cells progress into S phase without a concurrent loss of thermal resistance, using 45 degrees C induction doses even less toxic than used by other workers.  相似文献   

6.
The object of this study was to develop a map of G1 phase on the basis of the progressive changes taking place in the morphology of the prematurely condensed chromosomes as the cells traverse through G1 and then use this technique to determine the cell cycle location of normal and transformed cell populations in plateau phase. The morphology of the prematurely condensed chromosomes (PCC) of G1 cells in random populations was found to be highly variable. For a better understanding of the relationship between the morphology of the G1-PCC and their position within G1 phase, synchronized populations of Chinese hamster ovary (CHO) cells in early, mid-, and late G1 phase were fused with mitotic cells. Early G1 cells resulted in highly condensed G1-PCC, while late G1 cells gave very extended G1-PCC. Mid-G1 cells resulted in PCC of intermediate condensation. To test the validity of these criteria for mapping the position of a cell in the cell cycle, synchronous G1 cell populations were treated with a variety of metabolic inhibitors. Cycloheximide and actinomycin D were shown to block cell in early G1 phase, while excess thymidine and hydroxyurea blocked cells in early S phase. The results presented here indicate that, upon reaching plateau phase, normal cell populations (BALB-C mouse 3T3, human PA-2, and WI 38) stop in early G1, while most cells in transformed cell lines (CHO, HeLa, and mouse SV-3T3) accumulate in late G1.  相似文献   

7.
Heat treatment of CHO cells in the G1-phase of the cell cycle leads to chromatid-type aberrations in first posttreatment metaphases. Posttreatment of heat-treated cells with the alkylating agent trenimon leads to a synergistic effect on the production of chromatid-type exchanges. These results indicate that heat induces lesions which like the lesions produced by trenimon give rise to chromatid-type aberrations during the first posttreatment S-phase, and that these lesions can interact with each other to produce chromatid-type exchanges. Treatment of CHO cells in the G1-phase of the cell cycle with the restriction endonuclease Alu I induces chromosomal aberrations. Pretreatment of cells with heat leads to a reduction of Alu I induced chromosome-type aberrations. When cells are allowed to recover after heat treatment for 22 h, the aberration frequencies produced by Alu I are the same as in cells not treated with heat. These findings can be explained by assuming that heat-induced accumulation of accessory proteins in the chromatin protects the DNA from being cut by Alu I, and that the cells recovered from the heat-induced protein accumulation after 22 h.  相似文献   

8.
We have extended our studies on the cell cycle dependence of thermotolerance to include HeLa cells heated at 45.0 degrees C to compare the results to Chinese hamster ovary (CHO) cells. We found that asynchronous HeLa cells were more resistant to heat than CHO cells but showed a similar development and decay of thermotolerance. Flow cytometry (FCM) was used to study redistributions in the cell cycle after an initial heat dose. Cells heated for 35 min at 45.0 degrees C were delayed in G1 by about 7 h compared to controls, with delays in late S and G2/M phase also. The heat sensitivity varied through the cell cycle; G1 cells were the most resistant to heat, while S-phase cells were uniformly sensitive throughout S phase, and G2 cells were resistant. Thermotolerance could be induced and expressed in early or late S-phase cells, but to a lesser extent than for G1 cells. The results were similar in many respects to CHO cells, but there were significant differences.  相似文献   

9.
Endothelin-1 (ET-1) is a vasoactive peptide which signals through two G-protein coupled receptors, endothelin receptor A (ETA) and B (ETB). We determined that ET-1 activation of its ETB receptor in stably cDNA transfected CHO cells leads to a 55% reduction in cell number by end-point cell counting and a 35% decrease in cell growth by a real-time cell-substrate impedance-based assay after 24h of cell growth. When CHO ETB cells were synchronized in the late G1 cell cycle phase, ET-1 delayed their S phase progression compared to control by 30% as determined by [(3)H]-thymidine incorporation. On the other hand, no such delay was observed during late G2/M to G1 transit when cells were treated with ET-1 after release from mitotic arrest. Using the cell-substrate impedance-based assay, we observed that ET-1 induces opposing morphological changes in CHO ETA and CHO ETB cells with ETB causing an increase in the cell footprint and ETA a decrease. Likewise, in pulmonary artery smooth muscle cells, which express both ETA and ETB receptors, ET-1 induces an ETA-dependent contraction and an ETB dependent dilation. These results are shedding light on a possible beneficial role for ETB in diseases involving ET-1 dysfunction such as pulmonary hypertension.  相似文献   

10.
We report here for the first time that germanium oxide (GeO(2)) blocks cell progression. GeO(2) is not genotoxic to Chinese hamster ovary (CHO) cells and has limited cytotoxicity. However, GeO(2) arrests cells at G2/M phase. The proportion of cells stopped at G2/M phase increased dose-dependently up to 5 mM GeO(2) when treated for 12 h, but decreased at GeO(2) concentration was greater than 5 mM. Analysis of 5-bromodeoxyuridine-labeled cells indicated that GeO(2) delayed S phase progression in a dose-dependent manner, and blocked cells at G2/M phase. Microscopic examination confirmed that GeO(2) treatment arrested cells at G2 phase. Similar to several other events that cause G2 block, the GeO(2)-induced G2 block can also be ameliorated by caffeine in a dose- and time-dependent manner. To explore the mechanism of G2 arrest by GeO(2), cyclin content and cyclin-dependent kinase activity were examined. Cyclin B1 level was not affected after GeO(2) treatment in CHO cells. However, GeO(2) decreased p34(cdc2) kinase (Cdk1) activity. The kinase activity recovered within 9 h after GeO(2) removal and correlated with the transition of G2/M-G1 phase of the cells. This result suggests that GeO(2) treatment reduces Cdk1 activity and causing the G2 arrest in CHO cells.  相似文献   

11.
The effect of serum and temperature elevation on proliferation has been studied in synchronized mouse neuroblastoma (Neuro-2A) cells. The effects of serum were studied on the induction of (a) mitotic delay due to a non-lethal heat treatment (30 min at 42.7 degrees C) and (b) the loss of colony-forming capacity after a more extensive heat treatment (45 min at 44 degrees C or a continuous 42.7 degrees C heat treatment). The following results were obtained. Under conditions of serum depletion, cell cycle extension of heated G1 phase cells was more than that of heated G2 phase cells. Serum protected against heat-induced alterations of cell cycle progression in G1- but not in G2 phase cells. This effect of serum could be mimicked by a supplement to the medium of human transferrin, bovine pancreas insulin and selenium, and was correlated with protection of protein synthesis. Serum also affected heat-induced cell killing. Under conditions of serum depletion, G1 phase cells were more resistant to heat compared to G2 cells. The presence of serum during heat treatment further increased the thermoresistance of G1 phase cells, but did not affect sensitivity of G2 phase cells. This effect of serum could not be mimicked by a supplement of transferrin, insulin and selenium. These results indicate that serum protects G1 phase cells for heat-induced changes of cell cycle progression as well as on cell survival, but the mechanisms involved in both phenomena seem to be different.  相似文献   

12.
We studied mitogen-activated protein kinase (MAPK) activities during the cell cycle of Chinese hamster ovary (CHO) cells using site-specific antibodies against extracellular signal-regulated kinase-1, a 44-kDa MAPK (Boulton, T.G., Yancopoulos, G.D., Gregory, J.S., Slauer, C., Moomaw, C., Hsu, J., and Cobb, M.H. (1990) Science 249, 64-67). These antibodies detected two distinct MAPKs (44- and 42-kDa MAPKs) in CHO cells. CHO cells were arrested at metaphase in the M phase by treatment with nocodazole, and activities of MAPKs were analyzed at specific time points after release from arrest. Immune complex kinase assay and renaturation and phosphorylation assay in substrate-containing gel revealed that both 44- and 42-kDa MAPKs had activities in the G1 through S and G2/M phases and were activated biphasically, in the G1 phase and around the M phase. MAPKs were inactivated in metaphase-arrested cells. The amount of MAPKs did not change significantly in the cell cycle. In the G1, S, and G2/M phases, MAPKs were phosphorylated on both tyrosine and threonine residues and dephosphorylated in metaphase-arrested cells. Our data suggest that MAPKs may play some role in the cell cycle other than G0/G1 transition.  相似文献   

13.
Relatively little is known about the mechanisms used by somatic cells to regulate the replication of the centrosome complex. Centrosome doubling was studied in CHO cells by electron microscopy and immunofluorescence microscopy using human autoimmune anticentrosome antiserum, and by Northern blotting using the cDNA encoding portion of the centrosome autoantigen pericentriolar material (PCM)-1. Centrosome doubling could be dissociated from cycles of DNA synthesis and mitotic division by arresting cells at the G1/S boundary of the cell cycle using either hydroxyurea or aphidicolin. Immunofluorescence micros-copy using SPJ human autoimmune anticentrosome antiserum demonstrated that arrested cells were able to undergo numerous rounds of centrosome replication in the absence of cycles of DNA synthesis and mitosis. Northern blot analysis demonstrated that the synthesis and degradation of the mRNA encoding PCM-1 occurred in a cell cycle-dependent fashion in CHO cells with peak levels of PCM-1 mRNA being present in G1 and S phase cells before mRNA amounts dropped to undetectable levels in G2 and M phases. Conversely, cells arrested at the G1/S boundary of the cell cycle maintained PCM-1 mRNA at artificially elevated levels, providing a possible molecular mechanism for explaining the multiple rounds of centrosome replication that occurred in CHO cells during prolonged hydroxyurea-induced arrest. The capacity to replicate centrosomes could be abolished in hydroxyurea-arrested CHO cells by culturing the cells in dialyzed serum. However, the ability to replicate centrosomes and to synthesize PCM-1 mRNA could be re- initiated by adding EGF to the dialyzed serum. This experimental system should be useful for investigating the positive and negative molecular mechanisms used by somatic cells to regulate the replication of centrosomes and for studying and the methods used by somatic cells for coordinating centrosome duplication with other cell cycle progression events.  相似文献   

14.
Progression through the cell cycle is dependent upon numerous external factors (growth factors, extracellular matrix components) which exert their effects through the activation of signal transduction networks. During last years we have studied the regulation of progression through the ongoing CHO cell cycle. Recently, we have demonstrated that in CHO cells at least two serum dependent points exist in G1 phase that lead to different cellular responses. The first point is located immediately after mitosis and is suggested to link with apoptosis, while the second is located in late G1 phase and probably corresponds to the classical restriction point R. Because of the suggested link with apoptosis of the restriction point in early G1 phase, we have studied the possible role of PI 3-K in cell cycle progression through the ongoing G1 phase of CHO cells. In the presence of the PI 3-K inhibitors wortmannin or LY294002, cells were arrested during early G1 phase, leading to the expression of cleaved caspase-3, a central mediator of apoptosis. Addition of AP-2, an inhibitor of PKB, the downstream substrate of PI 3-K, at several time points during G1 phase demonstrated that inhibition during early G1 phase caused cell cycle arrest, while addition of the inhibitors during mid or late G1 phase had no effect on S phase entry. As for inhibition of PI 3-K, also inhibition of PKB resulted in expression of cleaved caspase-3. These results clearly demonstrate that a decision point exists in the early G1 phase of the cell cycle; in the presence of PKB activity the cells are continuing cell cycle progression, while in the absence of PKB activity the cells are induced for apoptosis.  相似文献   

15.
The ability of glycopeptides, isolated from bovine cerebral cortex, to alter cell division was studied by cell-cycle analyses. The results showed that glycopeptides arrested baby hamster kidney (BHK)-21 cells and Chinese hamster ovary (CHO) cells in the G2 phase of the cell cycle. Upon removal of the growth inhibition from arrested BHK-21 cells, the mitotic index in colchicine-treated cultures increased from 5 to 40% within 6 h and the increase in mitotic activity was accompanied by a complete doubling of all arrested cells within this 6- h time period. Determination of DNA content in growth-arrested BHK-21 cells showed that growth-arrested cells contained about twice the DNA of control cell cultures. Although CHO cells treated in a like manner with growth inhibitor could not be arrested for the same length of time as BHK-21 cells (18 h vs. 72 h before initiation of escape) and to the same degree (60% of the cell population vs. 99% of BHK-21 cells), the escape kinetics of CHO cells did indicate a G2 arrest. Approximately 3.5 h after escape began, CHO cell numbers in treated cultures attained the cell numbers found in control cultures. This rapid growth phase occurring in less than 4 h indicated that the growth inhibitor induced a G2 arrest-point in CHO cells that was not lethal since the entire arrested cell population divided.  相似文献   

16.
The metachromatic fluorochrome acridine orange was used to differentially stain DNA and RNA in Chinese hamster ovary (CHO) cells and in mitogen-stimulated human lymphocytes during their progression through the cell cycle. Green and red fluorescence of individual cells, representing cellular DNA and RNA, respectively, was measured by flow cytometry. CHO cells were synchronized by selective detachment at mitosis. Their rate of progression through G1 and subsequently through S phase correlated with the content of stainable RNA. The mean duration of the G1 phase was 5.2 hours for cells with high RNA content (highest 25 percentile population) and 8.1 hours for cells with low RNA (lowest 25 percentile). The duration of S phase was 5.9 and 7.5 hours for high- and low-RNA, 25 percentile subpopulations, respectively. Lymphocytes synchronized at the G1/S boundary by hydroxyurea or 5-fluorodeoxyuridine showed extremely high intercellular variation with respect to content of stainable RNA. After release from the block they traversed S phase at rates linearly proportional to the content of stainable RNA. The duration of S phase was five hours for cells with high RNA-, six to nine hours for cells with moderate RNA- and up to 27 hours for cells with minimal RNA-content. The data suggest that the rate of progression through the cell cycle of individual cells within a population may be correlated with the number of ribosomes per cell.  相似文献   

17.
To determine the effect of cell cycle position on protein synthesis, synchronized cell populations were metabolically labeled and the synthesis of the basic proteins, including histones, was examined by two-dimensional gel electrophoresis. Exponentially growing S49 mouse lymphoma or Chinese hamster ovary (CHO) cells were separated into G1 and S phase populations by centrifugal elutriation, selective mitotic detachment, fluorescence-activated cell sorting, or a combination of these, and pulse-labeled with radiolabeled amino acids. The histone proteins, both free and chromatin-bound, were completely resolved from some 300 other basic polypeptides in whole-cell lysates by a modification of the NEPHGE technique of O'Farrell, Goodman and O'Farrell (1977). Comparisons of matched autoradiograms from samples of G1 and S phase labeled cells revealed an equivalent rate of histone synthesis through the cell cycle of both S49 and CHO cells. Nuclei isolated from G1 phase S49 cells that were pulse-labeled contained between 13 and 15% of the newly synthesized nucleosomal histones present in S phase nuclei. Nuclei prepared from G1 phase cells that were pulse-labeled and then chased for 5 hr contained more than 90% of the labeled nucleosomal histones present in wholecell lysates. It therefore seems likely that differential alterations in the rate of histone synthesis do not occur to a significant degree as cells proceed through the cycle, but the association of newly synthesized histones with DNA takes place after the onset of DNA replication.  相似文献   

18.
Mammalian 8-oxo-2'-deoxyguanosine 5'-triphosphate pyrophosphohydrolases (8-oxo-dGTPases), such as MTH1, are believed to play the same antimutagenic role as their bacterial homologues, like MutT. Both decompose promutagenic 8-oxo-dGTP, a product of active oxygen's attack on dGTP. It is not known how 8-oxo-dGTPase expression and function are regulated. Therefore, we investigated the effect of cell population density, proliferation rate, and cell cycle phase on 8-oxo-dGTPase specific activity in cultured Chinese hamster ovary K1-BH4 (CHO) cells. With increasing cell population density (from 30 to 95% confluence), the activity of 8-oxo-dGTPase per milligram protein decreased by 33% (p =.007 by ANOVA) while cells shifted by 9% into the G(0)/G(1) phase, with a 5% drop in cells in S phase. Importantly, inhibition of the cells' proliferation rate by calf serum deprivation caused a more dramatic 23% shift toward the G(0)/G(1) phase and a 25% drop in S phase, but had no effect on 8-oxo-dGTPase activity. Likewise, no differences in the enzyme activity were observed within cell populations of different cell cycle phases separated by centrifugal elutriation. Thus, the present results exclude cell cycle-dependent regulation of 8-oxo-dGTPase activity in CHO cells or its simple dependence on proliferation rate. The observed decrease of 8-oxo-dGTPase activity with increasing cell population density might be related to augmentation of cell-to-cell contact.  相似文献   

19.
Tyrosyltubulin ligase (TTL) was found to be present in CHO and V79 Chinese hamster cells grown in tissue culture. The enzyme is soluble and requires potassium, magnesium, and ATP for maximum activity and requires tubulin as a substrate. TTL was analyzed through the cell cycle of V79 and CHO Chinese hamster cells. The enzyme showed two peaks of activity in V79 cells at 4 h and 7 h after mitotic selection, corresponding to the early S and mid to late S phases of the cell cycle. In CHO cells the enzyme displayed a major peak of activity at mid S and a minor peak or plateau during early S. Tubulin, as measured by (3H)colchicine binding, was shown to increase through S phase and reach a maximum late in the cycle during G2 approx. 3 h after maximum TTL activity.  相似文献   

20.
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