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.
The use of mitochondrial RNA as an indicator of apoptosis was investigated. Exposure of HA-1 fibroblastic cells to 10 H2O2 per 107 cells induced nuclear fragmentation, cell shrinkage, and internucleosomal DNA fragmentation, all characteristics of apoptosis. RNA extracted from control and apoptotic cultures, and analyzed by Northern blot hybridization, revealed a significant increase in the degradation of mitochondrial 16S ribosomal RNA (rRNA) that was associated with apoptosis. Conversely, minimal, if any, degradation of glyceraldehyde-3-phosphate dehydrogenase or actin mRNAs was observed. Similar results were obtained for HA-1 cells treated with the protein kinase inhibitor staurosporine, and for HT-2 T-lymphocytes induced to undergo apoptosis by interleukin-2 withdrawal. In addition, 16S rRNA degradation was an early event that was discernable well before chromatin condensation in hydrogen peroxide-treated HA-1 cells. These observations suggest that degradation of mitochondrial 16S ribosomal RNA is a new marker of mammalian cell apoptosis. © 1997 Elsevier Science Inc.  相似文献   

2.
Exponentially growing L929 cells were continuously exposed to 1 or 10 μMetoposide (VP-16). The effects of such treatment on cell growth, cycle distribution, morphology, and selected biochemical events were examined. DNA synthesis rates were markedly decreased and the protein/DNA ratio increased (unbalanced growth). Growth was blocked, with most cells being cycle arrested by 24 h in (late S–)G2–M. An asynchronous process of cell death then developed. Cells initially shrank into eosinophilic, trypan blue-excluding bodies, which were then released into the medium, and eventually became permeable to trypan blue. Transmission electron microscopy confirmed that dying cells acquired an apoptotic morphotype, with compaction and margination of chromatin, loss of microvilli, and shrinkage of cytoplasm and nucleus. Tissue transglutaminase activity and intensity of immunostaining rapidly increased in treated cultures. Internucleosomal DNA fragmentation could not be detected by agarose gel electrophoresis, yet flow cytometry revealed that the apoptotic bodies had a very low DNA fluorescence (≤10% of the 2nvalue). In agreement with the microscopic findings, this suggested that extensive DNA degradation had occurred in dead cells. While rates of cell loss from the monolayer amounted to 21 and 57% day−1(1 and 10 μMVP-16, respectively), apoptotic indexes largely underestimated the extent of the process. These indexes only measured the accumulation of apoptotic bodies, i.e., the balance between their generation and disposal. The latter occurred by mechanisms similar to those that operate in tissues: “secondary necrosis” or phagocytosis by viable homotypic cells in the monolayer (“homophagy”).  相似文献   

3.
The in vivo time course of the morphological changes and DNA degradation in castration-induced apoptotic prostate cells was studied from the earliest to the latest stage of the degeneration process. To study this problem, we first induced apoptotic prostate cells in rats by castration for 3 days and then promptly and continuously blocked the death of healthy prostatic cells in the castrated rats by in vivo testosterone replacement. Because testosterone replacement could not stop the irreversible lysis of already damaged prostate cells, apoptotic cells at different stages of the degeneration process were eliminated sequentially from the prostate after the healthy prostate cells had been protected. Prostate cells at the earliest stage of apoptosis at the time when the castrated rats received testosterone replacement disappeared last. By tracing the morphological and DNA degradation of apoptotic cells after hormone treatment, we estimated the time course of prostate cell death from the early to the final stage. In the morphological evolution of apoptotic prostate cells, the clumping of nuclear chromatin, the degeneration of cytoplasm and the involution of the cell surface occurred and progressed simultaneously, resulting in the rapid formation of apoptotic bodies that were gradually digested by other cells. The DNA ladders of apoptotic cells were progressively cleaved into a mononucleosomal subunit that was further degraded at an additional site, generating a heterogeneous population of small nucleotides. The final digestion of DNA fragments occurred within the apoptotic bodies. The whole course of prostate cell death after castration took about 44 h.  相似文献   

4.
He L  Kim BY  Kim KA  Kwon O  Kim SO  Bae EY  Lee MS  Kim MS  Jung M  Moon A  Bae K  Ahn JS 《Cellular signalling》2007,19(8):1713-1721
DNA damaging agents, such as camptothecin, and ionizing radiation (IR), can induce both NF-kappaB activation and apoptosis, however, the mechanism of their inter-regulation is not yet clear. In the present study, we discovered that Akt1 is degraded when cells deficient in Ataxia Telangiectasia mutated (ATM) were treated to CPT for apoptosis induction. While CPT-induced NF-kappaB activation could not be detected in ATM-deficient AT5BIVA cells, caspase-3 activation occurred and was even further enhanced by pretreatment with proteasome inhibitor-1 (Pro1), a NF-kappaB inhibitor. In contrast, activation of NF-kappaB but not of caspase-3 by CPT could be found in normal MRC5CV1 cells. NF-kappaB inhibition by Pro1, dominant negative mutant IkappaBalpha (S32/36) or p65 (N250), however, induced the caspase-3 activation in the normal cells, indicating the role of ATM-mediated NF-kappaB activation against cell apoptosis. On the other hand, interestingly, CPT significantly reduced the level of Akt1, this effect further enhanced by Pro1 pretreatment in AT5BIVA cells. In MRC5CV1 cells, however, Akt1 level could be reduced only when CPT and NF-kappaB inhibitors were co-treated to the cells, and this reversed by DEVD-cho treatment, demonstrating the caspase-3-mediated Akt1 degradation. Moreover, although MRC5CV1 cells were much more resistant to CPT compared with AT5BIVA, wortmannin and LY294002 significantly increased the chemosensitivity of MRC5CV1 cells to CPT. Given the accumulating evidences demonstrating Akt as a promising anticancer therapeutic target, all these results suggest that DNA damage induced apoptosis could be regulated by ATM-mediated NF-kappaB activation, and that Akt1 degradation be necessarily required for this apoptotic process.  相似文献   

5.
The determination of whether a cell dies by apoptosis as opposed to necrosis is usually best made on the basis of distinct structural changes in the chromatin. These changes include extensive condensation of the chromatin and DNA fragmentation. We have shown that the cytotoxic drug bleomycin (BLM) is able to cleave the DNA between the nucleosomes when it enters into the cell. If sufficient amounts of BLM are internalized, the nuclear morphological changes characteristic of apoptosis are detected. In this work, we describe the nuclear changes that occurred after DNA fragmentation as a function of the number of DNA double-strand breaks generated per cell and of the time after their generation. Our results show that DNA fragmentation and degradation of higher-order DNA structure were directly responsible for the nuclear morphological changes associated with apoptosis. During apoptosis reduced fluorescence with respect to the G0/G1 cell cycle region (the sub-G1 region) is often detected if fixed cells from cultures undergoing apoptosis are analyzed by flow cytometry. We demonstrate here that, depending on the extent of the DNA fragmentation and on ulterior changes in chromatin structure, the content of the fluorescent sub-G1 region can be either soluble pieces of DNA or apoptotic bodies or cells depleted in the DNA content by partial loss of fragmented DNA dissolved in the washing media and/or by the release of apoptotic bodies.  相似文献   

6.
Early indicators of apoptosis in mammalian cells are membrane potential breakdown (loss) in mitochondria (MPLM), chromatin condensation, DNA degradation, and phosphatidylserine exposure (PSE) on the outside plasma membrane. One aim of the present study was to determine the kinetics of these characteristics. These changes were measured by flow cytometry using the following methods: membrane potential of mitochondria was analysed using Mito Tracker Green and Red, PSE was analysed using annexin-V-FITC staining simultaneously with propidium iodide (PI) to detect membrane permeability; chromatin condensation was measured using the acid denaturation Acridine Orange (AO) method; DNA degradation was studied by the sub G1 method and the terminal transferase dUTP nick end-labelling (TUNEL) assay (labelling of strand breaks). HL-60 cells were induced to apoptosis by 3% ethanol and 1.5 microM camptothecin (CAM) and the kinetics of the apoptotic cells were measured. The same kinetics were found for chromatin condensation and DNA degradation indicating that these changes appeared at approximately the same time after induction. The MPLM and PSE kinetics showed a considerably later increase indicating that MPLM occurred downstream of DNA degradation and that plasma membrane changes occurred downstream of MPLM. The main aim of the study was to follow the fate of apoptotic cells after the appearance of the initial characteristics. The lifetime of apoptotic cells was studied by chase experiments. The inducing drug was removed after 4 h treatment and the disappearance of apoptoses recorded. An exponential decay was measured with a half life (T(1/2)) of 17.8 h. As a corollary from these experiments, camptothecin was found to induce apoptosis also in G1 and G2 phase cells, however, it took much longer to occur than in S phase cells. Using labelling of the plasma membrane with a fluorescent cell membrane linker, it was possible to show that the majority of apoptotic bodies as well as condensed apoptotic cells contain DNA and membrane. The degradation of these apoptotic bodies follows similar kinetics as those of the condensed apoptotic cells. The membrane remained considerably stable, there was no further loss in the next 7 days, after the first day when the apoptotic characteristics develop. It is concluded that the apoptosis programme is completed within a day and no further steps follow.  相似文献   

7.
8.
Caspase-dependent apoptosis is a controlled type of cell death characterized by oligonucleosomal DNA breakdown and major nuclear morphological alterations. Other kinds of cell death do not share these highly distinctive traits because caspase-activated DNase (DFF40/CAD) remains inactive. Here, we report that human glioblastoma multiforme-derived LN-18 cells do not hydrolyze DNA into oligonucleosomal fragments after apoptotic insult. Furthermore, their chromatin remains packaged into a single mass, with no signs of nuclear fragmentation. However, ultrastructural analysis reveals that nuclear disassembly occurs, although compacted chromatin does not localize into apoptotic nuclear bodies. Caspases become properly activated, and ICAD, the inhibitor of DFF40/CAD, is correctly processed. Using cell-free in vitro assays, we show that chromatin from isolated nuclei of LN-18 cells is suitable for hydrolysis into oligonuclesomal fragments by staurosporine-pretreated SH-SY5Y cytoplasms. However, staurosporine-pretreated LN-18 cytoplasms do not induce DNA laddering in isolated nuclei from either LN-18 or SH-SY5Y cells because LN-18 cells express lower amounts of DFF40/CAD. DFF40/CAD overexpression makes LN-18 cells fully competent to degrade their DNA into oligonucleosome-sized fragments, and yet they remain unable to arrange their chromatin into nuclear clumps after apoptotic insult. Indeed, isolated nuclei from LN-18 cells were resistant to undergoing apoptotic nuclear morphology in vitro. The use of LN-18 cells has uncovered a previously unsuspected cellular model, whereby a caspase-dependent chromatin package is DFF40/CAD-independent, and DFF40/CAD-mediated double-strand DNA fragmentation does not warrant the distribution of the chromatin into apoptotic nuclear bodies. The studies highlight a not-yet reported DFF40/CAD-independent mechanism driving conformational nuclear changes during caspase-dependent cell death.  相似文献   

9.
Barley aleurone cells undergo programmed cell death (PCD) when exposed to gibberellic acid (GA), but incubation in abscisic acid (ABA) prevent PCD. We tested the hypothesis that PCD in aleurone cells occurs by apoptosis, and show that the hallmark of apoptosis, namely DNA cleavage into 180 bp fragments, plasma membrane blebbing, and the formation of apoptotic bodies do not occur when aleurone cells die. We show that endogenous barley aleurone nucleases and nucleases present in enzymes used for protoplast preparation degrade aleurone DNA and that DNA degradation by these nucleases is rapid and can result in the formation of 180 bp DNA ladders. Methods are described that prevent DNA degradation during isolation from aleurone layers or protoplasts. Barley aleurone cells contain three nucleases whose activities are regulated by GA and ABA. CA induction and ABA repression of nuclease activities correlate with PCD in aleurone cells. Cells incubated in ABA remain alive and do not degrade their DNA, but living aleurone cells treated with GA accumulate nucleases and hydrolyze their nuclear DNA. We propose that barley nucleases play a role in DNA cleavage during aleurone PCD.  相似文献   

10.
11.
U937 human myeloid leukemia cells respond to mild treatment with hydrogen peroxide and hyperthermia by undergoing apoptosis, an active mode of cell suicide. Higher concentrations of hydrogen peroxide, or longer incubation at the hyperthermic temperature, change the mode of cell death from apoptosis to the passive necrosis. Stress treatments cause a severe drop in the intracellular NAD concentration. 3-Aminobenzamide (3-ABA), a specific inhibitor of poly(ADP-ribosyl) polymerase (PARP), a nuclear enzyme which is activated by breaks in DNA to catabolize intracellular NAD, is capable of relieving such a drop. This suggests that breaks in DNA have been induced by both oxidative stress and heat shock, thereby activating PARP. Upon stress, NAD concentration has a first initial sharp drop; then, for mild stress treatments, it recovers, just when apoptosis begins to be detectable (8 h of recovery). At 20 h, when the apoptotic ladder-like pattern of DNA is visible, NAD concentration has dropped again, probably because of a second PARP activation due to the extensive DNA degradation that accompanies apoptosis. The presence of 3-ABA, concomitantly with the preservation of the intracellular NAD content, reduces the extent of apoptosis upon oxidative stress and strongly enhances cell survival, thus suggesting a role for PARP in triggering stress-induced apoptosis. All apoptotic U937 cells have a reduced NAD content, independently of the inducing agent; however, upon treatments which do not cause immediate DNA breaks, the drop in NAD concentration occurs only after the apoptotic ladder is detectable and can be ascribed to the activation of PARP by the free ends of DNA formed during the endonucleolitic degradation. Moreover, in these instances the inhibition of PARP, although effective in blocking the drop in NAD concentration, has no effect on apoptosis, thus being only circumstantial.  相似文献   

12.
Microparticles are small membrane-bound vesicles that are released from apoptotic cells during blebbing. These particles contain DNA and RNA and display important functional activities, including immune system activation. Furthermore, nucleic acids inside the particle can be analyzed as biomarkers in a variety of disease states. To elucidate the nature of microparticle nucleic acids, DNA and RNA released in microparticles from the Jurkat T and HL-60 promyelocytic cell lines undergoing apoptosis in vitro were studied. Microparticles were isolated from culture media by differential centrifugation and characterized by flow cytometry and molecular approaches. In these particles, DNA showed laddering by gel electrophoresis and was present in a form that allowed direct binding by a monoclonal anti-DNA antibody, suggesting antigen accessibility even without fixation. Analysis of RNA by gel electrophoresis showed intact 18s and 28s ribosomal RNA bands, although lower molecular bands consistent with 28s ribosomal RNA degradation products were also present. Particles also contained messenger RNA as shown by RT-PCR amplification of sequences for β-actin and GAPDH. In addition, gel electrophoresis showed the presence of low molecular weight RNA in the size range of microRNA. Together, these results indicate that microparticles from apoptotic Jurkat and HL-60 cells contain diverse nucleic acid species, indicating translocation of both nuclear and cytoplasmic DNA and RNA as particle release occurs during death.  相似文献   

13.
Organisms exposed to reactive oxygen species, generated endogenously during respiration or by environmental conditions, undergo oxidative stress. Stress response can either repair the damage or activate one of the programmed cell death (PCD) mechanisms, for example apoptosis, and finally end in cell death. One striking characteristic, which accompanies apoptosis in both vertebrates and yeast, is a fragmentation of cellular DNA and mammalian apoptosis is often associated with degradation of different RNAs. We show that in yeast exposed to stimuli known to induce apoptosis, such as hydrogen peroxide, acetic acid, hyperosmotic stress and ageing, two large subunit ribosomal RNAs, 25S and 5.8S, became extensively degraded with accumulation of specific intermediates that differ slightly depending on cell death conditions. This process is most likely endonucleolytic, is correlated with stress response, and depends on the mitochondrial respiratory status: rRNA is less susceptible to degradation in respiring cells with functional defence against oxidative stress. In addition, RNA fragmentation is independent of two yeast apoptotic factors, metacaspase Yca1 and apoptosis-inducing factor Aif1, but it relies on the apoptotic chromatin condensation induced by histone H2B modifications. These data describe a novel phenotype for certain stress- and ageing-related PCD pathways in yeast.  相似文献   

14.
BACKGROUND: Much attention is being paid to the biologic effects of magnetic fields (MFs). Although MFs enhance tumorigenesis, they are neither mutagenic nor tumorigenic. The mechanism of their tumorigenic effect has not been elucidated. METHODS: To investigate the effect of MFs on apoptosis in HL-60 cells, we exposed the cells to static MFs of 6 mT generated by a magnetic disk of known intensity. Apoptosis was triggered by the DNA topoisomerase I inhibitor, camptothecin (CPT). Activation of caspases in situ using the fluorochrome-labeled inhibitor (FLICA) method and determination of plasma membrane integrity by excluding propidium iodide (PI) were measured by both laser scanning cytometry (LSC) and flow cytometry (FC). LSC and FC identified cells at three sequential stages of their demise: early apoptosis (cells with activated caspases and PI negative); late apoptosis (cells with activated caspases but unable to exclude PI); secondary necrosis (cells with apoptotic morphology no longer stained with FLICA, not excluding PI). RESULTS: MF alone did not induce any apoptogenic or necrogenic effect. CPT exposure led to the sequential appearance of apoptotic cells. In the presence of CPT and MF, the overall proportion of cells undergoing apoptosis was not significantly changed. However, we consistently observed a significant increase in the frequency of late apoptotic/necrotic cells when compared with samples treated with CPT alone (P < 0.001), as well as a decrease in the percentage of early apoptotic cells (P = 0.013). The data obtained by FC and LSC were consistent with each other, showing a similar phenomenon. CONCLUSION: Whereas MF alone or with CPT did not affect overall cell viability, it accelerated the rate of cell transition from apoptosis to secondary necrosis after induction of apoptosis by the DNA-damaging agent, CPT. Modulation of the kinetics of the transition from apoptosis to secondary necrosis by MF in vivo may play a role in inflammation and tumorigenesis.  相似文献   

15.
Discrete cleavages within 28S rRNA divergent domains have previously been found to coincide with DNA fragmentation during apoptosis. Here we show that rRNA and DNA cleavages can occur independently in apoptotic cells, i.e. that the previously observed correlation is likely to be coincidental. In HL-60 cells, apoptosis with massive DNA fragmentation could be induced without any signs of rRNA cleavage. The opposite situation; rRNA cleavage without concomitant internucleosomal DNA fragmentation, was found in okadaic acid-treated Molt-4 cells. Other leukemia cell lines underwent apoptosis either without (K562 and Molt-3) or with (U937) both forms of polynucleotide cleavage. In K562 cells transfected with a temperature-sensitive p53 mutant, internucleosomal DNA fragmentation but not 28S rRNA cleavage was inducible by wild-type p53 expression. The absence of apoptotic rRNA cleavage in some cell types suggests that this phenomenon is tightly regulated and unrelated to DNA fragmentation or a presumed scheme for general macromolecular degradation in apoptotic cells.  相似文献   

16.
17.
Cell nucleus and DNA fragmentation are not required for apoptosis   总被引:30,自引:3,他引:27       下载免费PDF全文
Apoptosis is the predominant form of cell death and occurs under a variety of physiological and pathological conditions. Cells undergoing apoptotic cell death reveal a characteristic sequence of cytological alterations including membrane blebbing and nuclear and cytoplasmic condensation. Activation of an endonuclease which cleaves genomic DNA into internucleosomal DNA fragments is considered to be the hallmark of apoptosis. However, no clear evidence exists that DNA degradation plays a primary and causative role in apoptotic cell death. Here we show that cells enucleated with cytochalasin B still undergo apoptosis induced either by treatment with menadione, an oxidant quinone compound, or by triggering APO-1/Fas, a cell surface molecule involved in physiological cell death. Incubation of enucleated cells with the agonistic monoclonal anti-APO-1 antibody revealed the key morphological features of apoptosis. Moreover, in non-enucleated cells inhibitors of endonuclease blocked DNA fragmentation, but not cell death induced by anti-APO-1. These data suggest that DNA degradation and nuclear signaling are not required for induction of apoptotic cell death.  相似文献   

18.
Castrated male rats were treated with the reversible S1-phase cell cycle blocking drug, mimosine, and the effects of this drug on prostate cell apoptosis was characterized. At a single dose of mimosine (25 mg/kg/day), we found that the internucleosomal DNA fragmentation associated with apoptosis was partially suppressed in the rat ventral prostate at all early time points (24, 48 and 72 h) analyzed post-castration. This suppression was dose-dependent, and treatment with mimosine up to 150 mg/kg/day was sufficient to reduce the internucleosomal DNA fragmentation in the prostate by 90% at 72 h post-castration. Intriguingly, this drug did not suppress the induction of mRNAs for several apoptosis-associated gene products in the ventral prostate gland (bcl-2, p53, TGF-beta and SGP-2/clusterin). Moreover, this treatment did not suppress the histological appearance of apoptotic bodies in the ventral prostate detectable by fast green staining of thin sections of tissue. The apoptotic bodies present in mimosine-treated regressing ventral prostate tissues, however, were refractory to labeling by the in situ gap labeling method, further demonstrating lack of nuclear DNA fragmentation in the condensed nuclei of apoptotic cells. In summary, the cell cycle-blocking drug mimosine does not appear to affect the rate of apoptosis in the regressing rat ventral prostate gland. However, this drug was capable of suppressing the nuclear DNA fragmentation associated with androgen-regulated prostate cell apoptosis. These results support the concept that nuclear DNA fragmentation is not obligatory for apoptosis. Additionally, they imply that cell cycle movement from the G1/S-phase boundary might be important for the terminal DNA degradation associated with androgen-regulated prostate cell apoptosis.  相似文献   

19.
In order to understand the role of actin microfilaments in the apoptotic process, we followed their evolution during tumor necrosis factor-alpha (TNF)-induced apoptosis in bovine aortic endothelial (BAE) cells. Using Western blotting analysis and immunofluorescence microscopy, we observed that the actin microfilaments network was disrupted in apoptotic cells. Depolymerization of F-actin was concomitant with internucleosomal DNA degradation and with the morphological changes associated with apoptotic cell death. However, using the actin microfilament disrupting agent, cytochalasin, we present evidence that the formation of blebs leading to apoptotic cell fragmentation requires neopolymerization of actin. Indeed, in the presence of cyochalasin, induction of apoptosis (internucleosomal DNA degradation) in BAE cells by TNF and cycloheximide was not associated with these classical morphological markers of apoptosis. Moreover, when added to BAE cells showing incipient apoptotic fragmentation, cytochalasin E reversed this process. We also observed an accumulation of actin at the basis of the apoptotic bodies in formation in these cells. Together, these results suggest that the actin network of flattened cells is disrupted concomitantly to the morphological modifications associated to the apoptotic cell death, and that the cytochalasin-sensitive reorganisation of actin is required to the formation of apoptotic blebs.  相似文献   

20.
BACKGROUND: Duration of apoptosis, from onset to final disintegration of the cell, is often short and variable. The apoptotic index (AI), as a snapshot of a transient event of variable length, does not truly represent incidence of apoptosis in the studied cell population. We recently proposed to estimate the cumulative apoptotic index (CAI) by inducing stathmo-apoptosis. A fluorescent inhibitor of caspases (FLICA) FAM-VAD-FMK is used to arrest the process of apoptosis and thereby prevent cell disintegration. Simultaneously, the arrested/apoptotic cells become FLICA-labeled. In the present study, this approach was applied to measure kinetics of HL-60 cell entrance into apoptosis induced via cell surface death receptor or a mitochondria-initiated pathway. Materials and Methods Cultures of HL-60 cells were treated with either TNF-alpha or camptothecin (CPT) in the absence or constant presence of 10-50 microM FLICA. The CAI was measured at different time points for up to 48 h by flow cytometry. Bivariate analysis of DNA content and cell labeling with FLICA was used to correlate apoptosis with the cell-cycle position. RESULTS: Selective loss of apoptotic cells seen in HL-60 cell cultures exposed to either TNF-alpha or CPT alone was prevented in cultures containing FLICA. Addition of FLICA alone had no effect on cell viability. The percentage of FLICA-labeled cells was plotted as a function of time after addition of TNF-alpha or CPT. The rate of cell entry to apoptosis was subsequently estimated from the slopes of the stathmo-apoptotic plot. The slopes revealed that the TNF-alpha or CPT-treated cells asynchronously underwent apoptosis with a stochastic-like kinetics and at two different rates. About 50% of cells in the TNF-alpha-treated cultures underwent apoptosis during the initial 6 h at a rate of approximately 8% of cells per hour; the remaining cells were undergoing apoptosis at a rate of approximately 2.5% of cells per hour for up to 24 h. Also, about 50% of the CPT-treated cells, predominantly those in S phase of the cell cycle, underwent apoptosis within the initial 8 h of CPT exposure, at a rate of approximately 7% of cells per hour. Remaining cells were undergoing apoptosis at a rate of approximately 1% of cells per hour during up to 48 h exposure to CPT. Spontaneous apoptosis in the untreated cultures occurred at a rate of 0.2% of cells per hour. CONCLUSIONS: This approach provides a means for measuring the kinetics of cell entrance to apoptosis (caspase activation) in large populations of cells in relation to the cell-cycle position.  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