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.
Increasing the knowledge of various cell cycle kinetic parameters, such as the length of the cell cycle and its different phases, is of considerable importance for several purposes including tumor diagnostics and treatment in clinical health care and a deepened understanding of tumor growth mechanisms. Of particular interest as a prognostic factor in different cancer forms is the S phase, during which DNA is replicated. In the present paper, we estimate the DNA replication rate and the S phase length from bromodeoxyuridine-DNA flow cytometry data. The mathematical analysis is based on a branching process model, paired with an assumed gamma distribution for the S phase duration, with which the DNA distribution of S phase cells can be expressed in terms of the DNA replication rate. Flow cytometry data typically contains rather large measurement variations, however, and we employ nonparametric deconvolution to estimate the underlying DNA distribution of S phase cells; an estimate of the DNA replication rate is then provided by this distribution and the mathematical model.  相似文献   

2.
The kinetics of DNA replication were analyzed in the second S phase following UV irradiation of Chinese hamster ovary cells synchronized at the beginning of S phase. The cells were synchronized by treating cells selected in mitosis with hydroxyurea for 9 h. Following UV irradiation, the cells were allowed to progress until the next mitosis; at which time they were resynchronized at the beginning of the second S phase by the same procedure. The kinetics of DNA replication were determined by measuring the proportion of DNA which achieved hybrid buoyant density on CsCl density gradients as a function of the time of incubation in the presence of 5-bromodeoxyuridine.The results of these experiments showed that even though the rate of DNA replication is substantially depressed during the first S phase following UV irradiation with a fluence of 5 J/m2, the rate has recovered to the extent that it is indistinguishable from the unirradiated control by the time the cells have entered their second S phase. It was concluded from these observations that the lesions in DNA which caused the rate of DNA replication to be initially depressed during the first S phase have been either removed or modified such that they no longer are able to cause a reduction in the rate of DNA replication in the second S phase following UV irradiation.  相似文献   

3.
When synchronous cells of the eucaryotic microorganism Chlorella sorokiniana growing in nitrate medium were challenged to synthesize an ammonium-inducible nicotinamide adenine dinucleotide phosphate-specific glutamate dehydrogenase (NADP-GDH) at frequent intervals during the cell cycle the initial rate of induction (i.e., enzyme potential) of this enzyme increased in an approximately linear manner until the period of DNA replication (i.e., S phase). During the S phase, NADP-GDH potential exhibited a positive rate change proportional to the step increase in DNA level. The timing of this rate change was insensitive to large changes in cellular growth rate. This rate change could be blocked within the first cell cycle by specific inhibition of DNA replication with 2'-deoxyadenosine. The approximately linear increase in NADP-GDH potential and also of total cellular protein observed before and after the S phase is proposed to be a result of the increasing photosynthetic capacity of the cell during the cell cycle.  相似文献   

4.
The time and coordination of cell cycle events were examined in the budding yeast Saccharomyces cerevisiae. Whole-cell autoradiographic techniques and time-lapse photography were used to measure the duration of the S, G1, and G2 phases, and the cell cycle positions of "start" and bud emergence, in cells whose growth rates were determined by the source of nitrogen. It was observed that the G1, S, and G2 phases underwent a proportional expansion with increasing cell cycle length, with the S phase occupying the middle half of the cell cycle. In each growth condition, start appeared to correspond to the G1 phase/S phase boundary. Bud emergence did not occur until mid S phase. These results show that the rate of transit through all phases of the cell cycle can vary considerably when cell cycle length changes. When cells growing at different rates were arrested in G1, the following synchronous S phase were of the duration expected from the length of S in each asynchronous population. Cells transferred from a poor nitrogen source to a good one after arrest in G1 went through the subsequent S phase at a rate characteristic of the better medium, indicating that cells are not committed in G1 to an S phase of a particular duration.  相似文献   

5.
Although the molecular enzymology of DNA replication is well characterised, how and why it occurs in discrete nuclear foci is unclear. Using fission yeast, we show that replication takes place in a limited number of replication foci, whose distribution changes with progression through S phase. These sites define replication factories which contain on average 14 replication forks. We show for the first time that entire foci are mobile, able both to fuse and re-segregate. These foci form distinguishable patterns during S phase, whose succession is reproducible, defining early-, mid- and late-S phase. In wild-type cells, this same temporal sequence can be detected in the presence of hydroxyurea (HU), despite the reduced rate of replication. In cells lacking the intra-S checkpoint kinase Cds1, replication factories dismantle on HU. Intriguingly, even in the absence of DNA damage, the replication foci in cds1 cells assume a novel distribution that is not present in wild-type cells, arguing that Cds1 kinase activity contributes to the spatio-temporal organisation of replication during normal cell growth.  相似文献   

6.
The organization of the mammalian S phase was studied in synchronized mouse embryo cells in terms of the spatial relationship between replication units whose synthesis is initiated at different times in S phase and the rate of assimilation of replication units into high molecular weight DNA strands.The formation of high molecular weight nascent DNA strands several replication units in length was analyzed by velocity sedimentation in alkaline sucrose gradients and by isopycnic centrifugation in alkaline Cs2SO4/CsCl gradients. Differential labeling with an isotopic and a density label shows that replication units synthesized at different stages of the S phase are not found within the same high molecular weight polynucleotide strand. It is thus concluded that replication units duplicated at different stages of the S phase are spatially organized in clusters along the mammalian genome.The rate of formation of high molecular weight nascent DNA strands is at least 4 to 8 times slower than that predicted from the spatial organization of replication units and the rate of chain growth within replication units. It is concluded that the process of joining of the completed nascent strands of adjacent replication units plays a major role in the rate of completion of high molecular weight strands.  相似文献   

7.
We have investigated components of DNA replication in a serum-induced S phase of primate CV-1 cells. Using DNA fiber autoradiography, we found a relative decrease in the frequency of initiation events in mid-S compared with early and late S phase. The other components of DNA replication measured by autoradiography—synchrony of initiation events, size of replication units, incidence of bidirectional replication, and the rate of replication fork movement—remained constant through S phase. When fork movement was measured by density gradient analysis of BUdR- and [3H]-thymidine-substituted DNA, it was also found to remain constant. These results show that most components of DNA replication are invariable through a serum-induced S phase. The changes in initiation frequency support the view that it may be critical in the regulation of ongoing replication.  相似文献   

8.
Activation (in the following referred to as firing) of replication origins is a continuous and irreversible process regulated by availability of DNA replication molecules and cyclin-dependent kinase activities, which are often altered in human cancers. The temporal, progressive origin firing throughout S phase appears as a characteristic replication profile, and computational models have been developed to describe this process. Although evidence from yeast to human indicates that a range of replication fork rates is observed experimentally in order to complete a timely S phase, those models incorporate velocities that are uniform across the genome. Taking advantage of the availability of replication profiles, chromosomal position and replication timing, here we investigated how fork rate may affect origin firing in budding yeast. Our analysis suggested that patterns of origin firing can be observed from a modulation of the fork rate that strongly correlates with origin density. Replication profiles of chromosomes with a low origin density were fitted with a variable fork rate, whereas for the ones with a high origin density a constant fork rate was appropriate. This indeed supports the previously reported correlation between inter-origin distance and fork rate changes. Intriguingly, the calculated correlation between fork rate and timing of origin firing allowed the estimation of firing efficiencies for the replication origins. This approach correctly retrieved origin efficiencies previously determined for chromosome VI and provided testable prediction for other chromosomal origins. Our results gain deeper insights into the temporal coordination of genome duplication, indicating that control of the replication fork rate is required for the timely origin firing during S phase.  相似文献   

9.
Both in yeast and in higher eukaryotes, genomic instability often ensues when the G1/S transition machinery is deregulated and cells are forced to enter S phase prematurely. This case of acquired mutability is particularly important, since a majority of genes mutated in human cancers encode factors that influence the G1/S transition. The precocious G1/S transition often results in a sub-optimal S phase. Moreover, the problems generated in such an S phase can escape detection by the cellular surveillance systems, allowing undeterred mitosis. This review focuses primarily on budding yeast data, where progress has been made in the past couple of years towards a mechanistic understanding of the underlying processes. A dual surveillance system is discussed, which relies on the presence of licensed but unfired origins and stalled replication forks to deter mitosis until replication is complete. Normally, this dual surveillance system allows S phase to be flexible in duration in a variety of growth conditions, when the fork density and/or fork progression rates can vary widely. However, precocious exit from G1 can have a disabling effect on this surveillance system. Premature exit from G1 can cut short the licensing of origins and the accumulation of resources for the upcoming replication, while giving a cell a false indication that it is metabolically ready to conduct S phase.  相似文献   

10.
Effects of temperature on the yeast cell cycle analyzed by flow cytometry   总被引:1,自引:0,他引:1  
M Vanoni  M Vai  G Frascotti 《Cytometry》1984,5(5):530-533
The effects of temperature (in the range 15-36 degrees C) on growth and the nuclear and budding cycle have been studied in populations of the yeast Saccharomyces cerevisiae exponentially growing in batch on yeast nitrogen base (YNB) glucose medium. The maximal rate of exponential growth is achieved at 30 degrees C, and a transition point is apparent at about 20 degrees C. At all tested temperatures DNA replication begins when cells are still unbudded and both the budded period and the postreplicative period have the same temperature dependence. A temperature compensatory mechanism seems to operate in S phase, during which duration remains relatively constant, in the range 21-36 degrees C, while duration of G2+ M phases shows a much more pronounced temperature dependence. The results are discussed in terms of a cell-cycle model for budding yeast.  相似文献   

11.
A stochastic model for interpreting BrdUrd DNA FCM-derived data is proposed. The model is based on branching processes and describes the progression of the DNA distribution of BrdUrd-labelled cells through the cell cycle. With the main focus on estimating the S phase duration and its variation, the DNA replication rate is modelled by a piecewise linear function, while assuming a gamma distribution for the S phase duration. Estimation of model parameters was carried out using maximum likelihood for data from two different cell lines. The results provided quite a good fit to the data, suggesting that stochastic models may be a valuable tool for analysing this kind of data.  相似文献   

12.
The rate of fork movement during DNA replication in mammalian cells   总被引:1,自引:1,他引:0  
Yu. B. Yurov 《Chromosoma》1979,74(3):347-353
DNA fiber autoradiography was used to measure the rate of replication fork progression along replication units in human diploid cells. The rate in different replication units differs very significantly and lies within the range 0.1 to 1.2 m/min. However, no significant changes were found in the rate of fork movement along single replication units operating during long intervals of S phase. Moreover, the fork progression rate is constant in many replication units of human cells.  相似文献   

13.
E36 ts24 is a temperature-sensitive cell cycle mutant which has been derived from the Chinese hamster lung cell line E36. This mutant is arrested in phase S when incubated at the restrictive temperature (40.3 degrees C) for growth. At this temperature, proliferation of the mutant cells ceases after 10 h. About 2 h earlier, DNA synthesis is arrested. These kinetic studies indicate that the execution point of the mutant cells is in early S phase well beyond the G1/S boundary. The pattern of replication bands in E36 ts24 cell grown for 9 h at 40.3 degrees C strengthen the kinetic studies and map the execution point to early S phase. The exact point of arrest of the mutant cells in phase S was mapped in early S phase near the execution point. At the point of arrest the cells continue to synthesize DNA at at a high rate but practically all of the newly synthesized DNA is degraded. This high rate of DNA degradation is limited to nascent DNA at the point of arrest. In the presence of 5-bromodeoxyuridine (5-BudR), the last E36 ts24 cells which reach mitosis at the restrictive temperature for growth show asymmetric replication bands which illustrate DNA degradation and resynthesis occurring in these cells at 40.3 degrees C.  相似文献   

14.
The discovery of G-rich oligonucleotides (GROs) that have non-antisense antiproliferative activity against a number of cancer cell lines has been recently described. This biological activity of GROs was found to be associated with their ability to form stable G-quartet-containing structures and their binding to a specific cellular protein, most likely nucleolin (Bates, P. J., Kahlon, J. B., Thomas, S. D., Trent, J. O., and Miller, D. M. (1999) J. Biol. Chem. 274, 26369-26377). In this report, we further investigate the novel mechanism of GRO activity by examining their effects on cell cycle progression and on nucleic acid and protein biosynthesis. Cell cycle analysis of several tumor cell lines showed that cells accumulate in S phase in response to treatment with an active GRO. Analysis of 5-bromodeoxyuridine incorporation by these cells indicated the absence of de novo DNA synthesis, suggesting an arrest of the cell cycle predominantly in S phase. At the same time point, RNA and protein synthesis were found to be ongoing, indicating that arrest of DNA replication is a primary event in GRO-mediated inhibition of proliferation. This specific blockade of DNA replication eventually resulted in altered cell morphology and induction of apoptosis. To characterize further GRO-mediated inhibition of DNA replication, we used an in vitro assay based on replication of SV40 DNA. GROs were found to be capable of inhibiting DNA replication in the in vitro assay, and this activity was correlated to their antiproliferative effects. Furthermore, the effect of GROs on DNA replication in this assay was related to their inhibition of SV40 large T antigen helicase activity. The data presented suggest that the antiproliferative activity of GROs is a direct result of their inhibition of DNA replication, which may result from modulation of a replicative helicase activity.  相似文献   

15.
J A D'Anna  R A Tobey 《Biochemistry》1989,28(7):2895-2902
Previous investigations showed that inhibition of DNA synthesis by hydroxyurea, aphidicolin, or 5-fluorodeoxyuridine produced large changes in the composition and nucleosome repeat lengths of bulk chromatin. Here we report results of investigations to determine whether the changes in nucleosome repeat lengths might be localized in the initiated replicons, as postulated [D'Anna, J. A., & Prentice, D. A. (1983) Biochemistry 22, 5631-5640]. In most experiments, Chinese hamster (line CHO) cells were synchronized in G1, or they were synchronized in early S phase by allowing G1 cells to enter S phase in medium containing 1 mM hydroxyurea or 5 micrograms mL-1 aphidicolin, a procedure believed to produce an accumulation of initiated replicons that arise from normally early replicating DNA. Measurements of nucleosome repeat lengths of bulk chromatin, the early replicating unexpressed metallothionein II (MTII) gene region, and a later replicating repeated sequence indicate that the changes in repeat lengths occur preferentially in the early replicating MTII gene region as G1 cells enter and become synchronized in early S phase. During that time, the MTII gene region is not replicated nor is there any evidence for induction of MTII messenger RNA. Thus, the results are consistent with the hypothesis that changes in chromatin structure occur preferentially in the early replicating (presumably initiated) replicons at initiation or that changes in chromatin structure can precede replication during inhibition of DNA synthesis. The shortened repeat lengths that precede MTII replication are, potentially, reversible, because they become elongated when the synchronized early S-phase cells are released to resume cell cycle progression.  相似文献   

16.
We measured the temporal order of replication of EcoRI segments from the murine immunoglobulin heavy-chain constant region (IgCH) gene cluster, including the joining (J) and diversity (D) loci and encompassing approximately 300 kilobases. The relative concentrations of EcoRI segments in bromouracil-labeled DNA that replicated during selected intervals of the S phase in Friend virus-transformed murine erythroleukemia (MEL) cells were measured. From these results, we calculated the nuclear DNA content (C value; the haploid DNA content of a cell in the G1 phase of the cell cycle) at the time each segment replicated during the S phase. We observed that IgCH genes replicate in the following order: alpha, epsilon, gamma 2a, gamma 2b, gamma 1, gamma 3, delta, and mu, followed by the J and D segments. The C value at which each segment replicates increased as a linear function of its distance from C alpha. The average rate of DNA replication in the IgCH gene cluster was determined from these data to be 1.7 to 1.9 kilobases/min, similar to the rate measured for mammalian replicons by autoradiography and electron microscopy (for a review, see H. J. Edenberg and J. A. Huberman, Annu. Rev. Genet. 9:245-284, 1975, and R. G. Martin, Adv. Cancer Res. 34:1-55, 1981). Similar results were obtained with other murine non-B cell lines, including a fibroblast cell line (L60T) and a hepatoma cell line (Hepa 1.6). In contrast, we observed that IgCh segments in a B-cell plasmacytoma (MPC11) and two Abelson murine leukemia virus-transformed pre-B cell lines (22D6 and 300-19O) replicated as early as (300-19P) or earlier than (MPC11 and 22D6) C alpha in MEL cells. Unlike MEL cells, however, all of the IgCH segments in a given B cell line replicated at very similar times during the S phase, so that a temporal directionality in the replication of the IgCH gene cluster was not apparent from these data. These results provide evidence that in murine non-B cells the IgCH, J, and D loci are part of a single replicon.  相似文献   

17.
DNA fiber autoradiography was used to measure the rate of replication fork movement and the size of replication units as a function of time during the S phase of synchronized Chinese hamster ovary cells. The rate of fork movement increased by about threefold from early S to later S phase, with the most dramatic change occurring in the first hour of S phase. On the other hand, the size of replication units did not vary significantly during S phase.  相似文献   

18.

Background

Eukaryotic cells seem unable to monitor replication completion during normal S phase, yet must ensure a reliable replication completion time. This is an acute problem in early Xenopus embryos since DNA replication origins are located and activated stochastically, leading to the random completion problem. DNA combing, kinetic modelling and other studies using Xenopus egg extracts have suggested that potential origins are much more abundant than actual initiation events and that the time-dependent rate of initiation, I(t), markedly increases through S phase to ensure the rapid completion of unreplicated gaps and a narrow distribution of completion times. However, the molecular mechanism that underlies this increase has remained obscure.

Methodology/Principal Findings

Using both previous and novel DNA combing data we have confirmed that I(t) increases through S phase but have also established that it progressively decreases before the end of S phase. To explore plausible biochemical scenarios that might explain these features, we have performed comparisons between numerical simulations and DNA combing data. Several simple models were tested: i) recycling of a limiting replication fork component from completed replicons; ii) time-dependent increase in origin efficiency; iii) time-dependent increase in availability of an initially limiting factor, e.g. by nuclear import. None of these potential mechanisms could on its own account for the data. We propose a model that combines time-dependent changes in availability of a replication factor and a fork-density dependent affinity of this factor for potential origins. This novel model quantitatively and robustly accounted for the observed changes in initiation rate and fork density.

Conclusions/Significance

This work provides a refined temporal profile of replication initiation rates and a robust, dynamic model that quantitatively explains replication origin usage during early embryonic S phase. These results have significant implications for the organisation of replication origins in higher eukaryotes.  相似文献   

19.
J A D'Anna  R A Tobey 《Biochemistry》1984,23(21):5024-5029
We have measured changes in histone H1 content and changes in chromatin structure of Chinese hamster (line CHO) cells blocked in early S phase by sequential use of isoleucine deprivation and blockade with 5-fluorodeoxyuridine or aphidicolin. Both the H1:core histone ratio in isolated nuclei and the H1 content of the cell are reduced 20-60%, depending on the duration of the block. The new deoxyribonucleic acid (DNA) synthesized during S-phase block has a shorter nucleosome repeat length than that of bulk chromatin, but it is nearly equally resistant as bulk DNA to attack by micrococcal nuclease. During the time that H1 content is decreasing, bulk chromatin also undergoes structural changes so that its nucleosome cores appear to be more closely packed along the DNA chain. The losses in H1 content and changes in chromatin structure are similar to those reported for cells blocked in early S phase by hydroxyurea [D'Anna, J. A., & Prentice, D. A. (1983) Biochemistry 22, 5631-5640]. The results suggest that losses of H1 and changes in chromatin structure are general events which occur when the elongation of initiated replicons or the joining of intermediate-sized DNA fragments is retarded during replication. They are consistent with the notions that H1 is lost from initiated replicons and/or the loss of H1 is part of an alarm response in the cell which might facilitate events leading to gene amplification.  相似文献   

20.
Quiescent cells reside in G0 phase, which is characterized by the absence of cell growth and proliferation. These cells remain viable and re-enter the cell cycle when prompted by appropriate signals. Using a budding yeast model of cellular quiescence, we investigated the program that initiated DNA replication when these G0 cells resumed growth. Quiescent cells contained very low levels of replication initiation factors, and their entry into S phase was delayed until these factors were re-synthesized. A longer S phase in these cells correlated with the activation of fewer origins of replication compared to G1 cells. The chromatin structure around inactive origins in G0 cells showed increased H3 occupancy and decreased nucleosome positioning compared to the same origins in G1 cells, inhibiting the origin binding of the Mcm4 subunit of the MCM licensing factor. Thus, quiescent yeast cells are under-licensed during their re-entry into S phase.  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