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 effects of butyrate upon the extents of phosphorylation of histones H1 and H1(0) during cell-cycle progression have been investigated. Chinese hamster (line CHO) cells were synchronized in early S phase and released into medium containing 0 or 15 mM butyrate to resume cell-cycle traverse into G1 of the next cell cycle. Cells were also mechanically selected from monolayer cultures grown in the presence of colcemid and 0 or 15 mM butyrate to obtain greater than 98% pure populations of metaphase cells. Although cell cycle progression is altered by butyrate, electrophoretic patterns of histones H1, H1(0), H3, and H4 indicate that butyrate has little, if any, effect on the extents of H1 and H1(0) phosphorylation during the cell cycle or the mitotic-specific phosphorylation of histone H3. Butyrate does, however, inhibit removal of extraordinary levels of histone H4 acetylation (hyperacetylation) during metaphase, and it appears to cause an increase in the content of H1(0) in chromatin during the S or G2 phases of the cell cycle.  相似文献   

2.
Histone proteins, which were assembled into chromatin using the Xenopus oocyte S-150 extract, were analyzed on acid-urea gels and Triton-acid-urea gels to determine their state of modification. We find that histone H4, which is present in a diacetylated form in the oocyte S-150, gradually loses its acetate groups as the DNA is packaged into chromatin. Thus, this process parallels the one observed in vivo during chromatin formation in growing eucaryotic cells. Histone H4 deacetylation in the oocyte S-150 is a DNA-dependent reaction. This reaction is blocked when butyrate (an inhibitor of histone deacetylase) is added at the onset of the chromatin assembly process. When butyrate is added at the end of the assembly process, no de novo acetylation of the nucleosomal histone H4 is observed. Chromatin with regularly spaced nucleosomes, displaying periodicities ranging from 160 to 220 base pairs, can be assembled in vitro with the oocyte S-150 (Rodríguez-Campos, A., Shimamura, A., and Worcel, A. (1989) J. Mol. Biol., in press). This chromatin may contain either deacetylated histone H4 when assembled under standard conditions or diacetylated H4 when assembled in the presence of butyrate. Both types of chromatin display identical structures upon digestion with nucleases. The potential applications of this system toward the study of the naturally occurring diacetylated histone H4 are discussed.  相似文献   

3.
The effects of sodium butyrate on [3H]thymidine incorporation and cell growth characteristics in randomly growing and synchronized HeLa S3 cells have been examined in an attempt to determine what effects, if any, butyrate has on S phase cells. Whereas 5 mM sodium butyrate rapidly inhibits [5H]thymidine incorporation in a randomly growing cell populations, it has no effect on incorporation during the S phase in cells synchronized by double thymidine block techniques. This lack of effect does not result from an impaired ability of the S phase cells to take up butyrate, since butyrate administration during this period leads to histone hyperacetylation that is identical with that seen with butyrate treatment of randomly growing cells. Furthermore, the ability to induce such hyperacetylation with butyrate during an apparently normal progression through S phase indicates that histone hyperacetylation probably has no effect on the overall process of DNA replication. Temporal patterns of [3H]thymidine incorporation and cell growth following release from a 24-h exposure to butyrate confirm blockage of cell growth in the G1 phase of the cell cycle. Thus, the inhibition by butyrate of [3H]thymidine incorporation in randomly growing HeLa S3 cell populations can be accounted for solely on the basis of a G1 phase block, with no inhibitory effects on cells already engaged in DNA synthesis or cells beyond the G1 phase block at the time of butyrate administration.  相似文献   

4.
5.
Sodium butyrate in millimolar concentrations causes an accumulation of acetylated histone species in a variety of vertebrate cell lines. In all lines tested, butyrate caused hyperacetylation of H3 and H4, and in rat IRC8 cells, H2A and H2B were also affected. In Friend erythroleukemic cells, butyrate also induces the synthesis of a nonhistone chromosomal protein, IP25. Butyrate does not affect the rate of histone acetylation in cell-free extracts or nuclei of Friend cells. Rather, this fatty acid inhibits histone deacetylation. Cell-free extracts of either control cells or butyrate-grown cells contain comparable levels of histone-deacetylating activity. This in vitro activity is inhibited by the addition of butyrate to the extracts. Thus butyrate appears to be an inhibitor of histone deacetylases both in vivo and in vitro.  相似文献   

6.
The short fatty acid, butyrate, which is produced by intestinal anaerobic bacteria in the colon, has inhibitory activity on histone deacetylases (HDACs). Treatment of the human colon cancer cell line, LS174T, with 1-2 mM sodium butyrate stimulated MUC2 mucin production, as determined by histological PAS staining of carbohydrate chains of mucin, and confirmed at the protein and mRNA levels by immunoblotting with anti-MUC2 antibody and real-time RT-PCR, respectively. Increases in acetylated histone H3 in the LS174T cells treated with butyrate suggest inhibition of HDACs in these cells. Butyrate-stimulated MUC2 production in the LS174T cells was inhibited by the MEK inhibitor, U0126, implicating the involvement of extracellular signal-regulated kinase (ERK) cascades in this process. Proliferation of the LS174T cells was inhibited by butyrate treatment. Although apoptotic nuclear DNA fragmentation could not be detected, cell-cycle arrest at the G0/G1 phase in the butyrate-treated cells was demonstrated by flow cytometry. Thus butyrate, an HDAC inhibitor, inhibits proliferation of LS174T cells but stimulates MUC2 production in individual cells.  相似文献   

7.
Cells were synchronized in G1 by isoleucine deprivation and then released into medium containing 1 mM hydroxyurea (HU), 5 micrograms mL-1 aphidicolin (APC), or 1 microgram mL-1 5-fluorodeoxyuridine (fl5dU). Coulter volume, content of histone H1 per unit DNA, turnover of histone H1, the extent of DNA elongation, and the survival of cells were measured as functions of time after release into the presence of the drugs. At the concentrations used in the experiments, the drug differ in their toxicity (fl5dU greater than HU greater than APC), induction of unbalanced cell growth, and the distribution of new DNA fragment sizes allowed during block, but they all (1) allow cells to enter S phase, (2) cause similar time-dependent losses of histone H1 per unit DNA, which begin as synchronized G1 cells begin to enter S phase, (3) retard DNA elongation beyond replicon size, and (4) retard the turnover of histone H1. The results indicate that loss of histone H1, inhibition of histone turnover, the retarded ligation of newly replicated DNA into bulk chromatin, and chromatin structural changes may be part of the cell's general response to inhibition of DNA replication. Since transient S phase block increases the frequencies of gene amplification [Mariani, B. D., & Schimke, R. T. (1984) J. Biol. Chem. 259, 1901-1910] and sister chromatid exchanges (SCE) [Rainaldi, G., Sessa, M. R., & Mariani, T. (1984) Chromosoma 90, 46-49], the observed changes in H1 content and chromatin organization may also be essential features of gene amplification and SCE.  相似文献   

8.
Antibodies specific for acetylated isoforms of histone H4 have been used to compare acetylation of this histone in interphase and metaphase cells. Two rabbit antisera (R5 and R6) were used, each specific for H4 molecules acetylated at one of the four possible acetylation sites, namely Lys-5 (R6) and Lys-12 (R5). Both antisera bound preferentially to the more-acetylated H4 isoforms (H4Ac2-4). To test for continued H4 acetylation in metaphase chromosomes. Chinese hamster ovary cells were blocked in metaphase and treated for one hour with the deacetylase inhibitor sodium butyrate. Isolated chromosomes were assayed for H4 acetylation by antibody labeling and flow cytometry. H4 acetylation was increased several fold by this brief butyrate treatment. The increase was in direct proportion to DNA content, with no evidence for exceptionally high- or low-labeling chromosomes. The results demonstrate that a cycle of H4 acetylation and deacetylation continues within metaphase chromosomes. Immunofluorescence microscopy showed labeling to be distributed throughout the chromosome, but with variable intensity. Western blotting and immunostaining with R5 and R6 showed a net reduction in labeling of H4 from metaphase cells, with major reductions in the more-acetylated isoforms H4Ac3-4. In contrast, labeling of H4Ac1 was reduced to a lesser extent (R6) or increased (R5). This increase indicates more frequent use of the acetylation site at lysine 12 in H4Ac1 from metaphase cells.  相似文献   

9.
10.
We have studied the effect of butyrate and other short-chain fatty acids on thyroid hormone nuclear receptors in C6 cells, a rat glioma cell line. Exposure of C6 cells to butyrate leads to increased levels of L-triiodothyronine (T3) in the nuclear and extranuclear compartments. The rise in nuclear binding is not merely a reflection of the higher cellular hormone content, and Scatchard analysis of T3 binding to isolated nuclei reveals that butyrate increases receptor number without changing affinity. The effect on the receptor is quantitatively important: a 48-h incubation with 2 mM butyrate increases nuclear binding by 2-3-fold, and 5 mM butyrate by 3-5-fold. Other short-chain fatty acids were found to similarly influence both nuclear receptor and extranuclear T3 levels with the following potency: butyrate greater than valerate greater than propionate greater than acetate. On the contrary, ketone bodies were ineffective. Butyrate increases receptor levels by decreasing receptor degradation, since the apparent t1/2 of receptor disappearance increased by approximately 3-fold in cells incubated with 2 mM butyrate for 48 h. The regulation of receptor number might be secondary to an action of butyrate on regions of the chromatin to which the receptor associates. We then examined the effect of butyrate on histone acetylation. The fatty acid had little effect in increasing the level of multiacetylated forms of H3 and H4 histone when studied in acid-urea gels, but it markedly inhibited the turnover of [3H] acetate from the histone fraction. There was a striking similarity in the dose-response of butyrate for increasing receptor levels and inhibiting histone deacetylation. Furthermore, a very close correlation between receptor levels and [3H]acetate release was also found when different short-chain fatty acids were used. We thus conclude that the effect of butyrate on the receptor could be explained by a modification of the chromatin structure of C6 cells secondary to acetylation.  相似文献   

11.
Dynamic histone acetylation of alfalfa (Medicago sativa) was studied in suspension cultures by short-term labeling with radioactive acetate. The relative labeling rates for the acetylated histones were in order of decreasing incorporation; H3.2 greater than H3.1 greater than H4 greater than H2B.1 greater than H2A.3. Histone H3 showed at least seven sites of acetylation, histone H2B.1 had six sites and histone H4 had five sites. Low numbers of acetylation sites were observed for histone H2B.2 and all histone H2A variants. The mass ratio, steady state acetylation and dynamic acetylation between major variant H3.1 and minor variant H3.2 were approx. 2:1, 1:2 and 2:5, respectively. Treatment of alfalfa cells with 50 mM n-butyrate did not lead to histone hyperacetylation, but instead interfered with histone acetylation labeling by acetate. The extent of apparent inhibition increased with time and concentration of butyrate. It is likely that the conversion of butyrate to acetylCoA results in dilution of the specific radioactivity of [3H]acetate in the acetylCoA pool thereby inhibiting the labeling reaction. This interpretation is supported by 14C-labeling of alfalfa acetylated histones by [1-14C]butyrate.  相似文献   

12.
H1 histone subtype genes differ in their expression patterns during the different stages of the cell cycle interphase. While the group of replication-dependent H1 histone subtypes is synthesized during S phase, the replacement histone subtype H1.0 is also expressed replication-independently in non-proliferating cells. The present study is the first report about the analysis of the cell cycle-dependent expression of all five replication-dependent H1 subtypes, the replacement histone H1.0 and the ubiquitously expressed subtype H1x. The expression of these H1 histone subtypes in HeLa cells was analysed on mRNA level by quantitative real-time RT-PCR as well as on protein level by immunoblotting. We found that after arrest of HeLa cells in G1 phase by treatment with sodium butyrate, the mRNA levels of all replication-dependently expressed H1 subtypes decreased, but to very different extent. During S phase the individual replication-dependently expressed H1 subtypes show similar kinetics regarding their mRNA levels. However, the variations in their protein amounts partially differ from the respective RNA levels which especially applies to histone H1.3. In contrast, the mRNA as well as the protein level of H1x remained nearly unchanged in G1 as well as during S phase progression. The results of the present study demonstrate that the cell cycle-dependent mRNA and protein expression of various H1 subtypes is differentially regulated, supporting the hypothesis of a functional heterogeneity.  相似文献   

13.
K562 is an established human erythroleukemia cell line, inducible for hemoglobin synthesis by a variety of compounds including n-butyrate. To elucidate the role of butyrate-induced histone acetylation in the regulation of gene expression in K562 cells, we isolated 20 variants resistant to the growth inhibitory effect of butyrate. Four variants having different degrees of resistance were selected for detailed study. All four were found to be resistant to the hemoglobin-inducing effect of butyrate, suggesting that the two aspects of butyrate response, restriction of growth and induction of hemoglobin synthesis, are coupled. Further, after (5 days) culture with butyrate, two of the four variants exhibit less acetylation of H3 and H4 histones than does the butyrate-treated parent. Analysis of histone deacetylases from the variants indicated that each variant was distinct and that butyrate resistance may be accounted for by decreased affinity of the variant enzymes for butyrate, increased affinity of the enzymes for acetylated histone, or both. The fact that variants selected for resistance to growth inhibition by butyrate are also deficient in butyrate-induced hemoglobin synthesis and have abnormal histone deacetylase activity argues for butyrate inducing K562 cells to synthesize hemoglobin and restrict growth via histone acetylation.  相似文献   

14.
Nuclei from butyrate-treated murine lymphosarcoma cells were incubated with different amounts of the polyanion heparin, which is known to interact predominantly with chromatin-associated histones. Unlike isolated histone H1, histone H1 in the nuclei of butyrate-treated cells was found to display an enhanced affinity for the binding to heparin as compared to histone H1 from control cells. Dephosphorylation of histone H1 as a result of butyrate treatment of the cells is discussed as a possible factor involved in the observed higher affinity of the protein for heparin.  相似文献   

15.
Linda Sealy  Roger Chalkley 《Cell》1978,14(1):115-121
The hyperacetylation of histones due to treatment of cultured cells with sodium butyrate has been studied. The hyperacetylation is due to inhibition of histone deacetylase. Other short chain fatty acids including acetic, isobutyric and propionic acid also produce increased modification. Histone H4 already deposited on the chromosome can be rapidly acetylated to the extent of about 70%. That 80% of histone H4 is acetylated after a 24 hr exposure to butyrate is due to the fact that incoming H4 histone is 100% acetylated and does not return to the parental unmodified form in the presence of butyrate.  相似文献   

16.
Structure of polyubiquitinated histone H2A   总被引:6,自引:0,他引:6  
B E Nickel  J R Davie 《Biochemistry》1989,28(3):964-968
We have recently demonstrated that trout liver histones H2A, H2B, and H2A.Z can be polyubiquitinated [Davie, J.R., Delcuve, G.P., Nickel, B.E., Moyer, R., & Bailey, G. (1987) Cancer Res. 47, 5407-5410]. In the present study we determined the arrangement of the ubiquitin molecules in polyubiquitinated histone H2A. Trout liver chromatin fragments. which had histone H1 removed, were digested with Staphylococcus aureus (V8 strain) protease which cleaves specifically on the carboxyl side of glutamic acid residues under the conditions used. The V8 protease readily degraded histone H2A and ubiquitinated (u) H2A at equivalent rates. One site in H2A and uH2A, the peptide bond between Glu 121 and Lys 122, was cleaved, yielding protein species cH2A and cuH2A, respectively. None of the other nucleosomal histones (H2B, H2A.Z, H3, and H4) including uH2B and uH2A.Z were sensitive to digestion. Trout liver histones cleaved with either V8 protease, histone H2A specific protease, or cyanogen bromide were resolved by two-dimensional gel electrophoresis and ubiquitinated peptides detected with anti-ubiquitin IgG. The results suggest that the major arrangement of ubiquitin in polyubiquitinated H2A is a chain of ubiquitin molecules joined to each other by isopeptide bonds to a ubiquitin molecule that is attached to the epsilon-amino group of lysine 119 of histone H2A.  相似文献   

17.
Nuclei from hepatoma tissue culture (HTC) cells were isolated by standard methods and incubated in media commonly used for nuclease digestions (DNAase I and micrococcal nuclease) and for in vitro RNA synthesis. During the incubation, histones can be deacetylated from both control cells and cells treated with 6 mM sodium butyrate to enhance the levels of histone acetylation. Deacetylation of histone is much more apparent in nuclei isolated from sodium butyrate-treated cells. Inclusion of 6 mM sodium butyrate in the incubation medium effectively inhibits the endogenous deacetylase activity acting on histones H3 and H4, whereas sodium acetate at the same concentration has very little inhibitory effect.  相似文献   

18.
The histone deacetylase inhibitor and potential anti-cancer drug sodium butyrate is a general inducer of growth arrest, differentiation, and in certain cell types, apoptosis. In human CCRF-CEM, acute T lymphoblastic leukemia cells, butyrate, and other histone deacetylase inhibitors caused G2/M cell cycle arrest as well as apoptotic cell death. Forced G0/G1 arrest by tetracycline-regulated expression of transgenic p16/INK4A protected the cells from butyrate-induced cell death without affecting the extent of histone hyperacetylation, suggesting that the latter may be necessary, but not sufficient, for cell death induction. Nuclear apoptosis, but not G2/M arrest, was delayed but not prevented by the tripeptide broad-range caspase inhibitor benzyloxycarbonyl-Val-Ala-Asp.fluoromethylketone (zVAD) and, to a lesser extent, by the tetrapeptide 'effector caspase' inhibitors benzyloxycarbonyl-Asp-Glu-Val-Asp.fluoromethylketone (DEVD) and benzyloxycarbonyl-Val-Glu-Ile-Asp.fluoromethyl-ketone (VEID); however, the viral protein inhibitor of 'inducer caspases', crmA, had no effect. Bcl-2 overexpression partially protected stably transfected CCRF-CEM sublines from butyrate-induced apoptosis, but showed no effect on butyrate-induced growth inhibition, further distinguishing these two butyrate effects. c-myc, constitutively expressed in CCRF-CEM cells, was down-regulated by butyrate, but this was not causative for cell death. On the contrary, tetracycline-induced transgenic c-myc sensitized stably transfected CCRF-CEM derivatives to butyrate-induced cell death.  相似文献   

19.
To approach experimentally changes of chromatin structure introduced by glucocorticoids, the histone H1 compositions of hormone-treated and non-treated mouse mammary tumor cells of the GR line [Ringold, G., Lasfargues, E. Y., Bishop, J. M. and Varmus, H. E. (1975) Virology 65, 135-147] were compared. To define the biologically important hormone concentration range, the cells were exposed to different concentrations of triamcinolone, a synthetic glucocorticoid. The induction of mouse mammary tumor virus (MMTV) RNA was measured by cDNA excess hybridization, and the amount of hormone bound to nuclei was determined by a filter-binding assay. Between 0.3 nM and 30 nM triamcinolone the relative increase in nuclear bound hormone corresponded well with the relative induction of MMTV RNA. The half-life of triamcinolone in nuclei of growing cells was 1 h, as measured by a pulse-chase experiment. Reversed-phase high-performance liquid chromatography of histone H1 resulted in its separation into four subfractions. The treatment of cells with biologically active glucocorticoid, 3 nM or 30 nM triamcinolone or 1 microM dexamethasone, resulted in changes in the relative amounts of two subfractions and to a positional shift of two subfractions as compared to untreated cells. No changes were observed after exposure to 3 nM dexamethasone, a concentration which does not induce MMTV RNA [Ringold, G. M., Yamamoto, K. R., Tomkins, G. M., Bishop, J. M. and Varmus, H. E. (1975) Cell 6, 299-305].  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