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 15 毫秒
1.
Effects of sodium butyrate, a new pharmacological agent, on cells in culture   总被引:32,自引:0,他引:32  
Sodium butyrate, at millimolar concentrations, when added to cell cultures produces many morphological and biochemical modifications in a reversible manner. Some of them occur in all cell lines. They concern regulatory mechanisms of gene expression and cell growth: an hyperacetylation of histone resulting from an inhibition of histone deacetylase and an arrest of cell proliferation are almost constantly observed. Some other modifications vary from one cell type to another: induction of proteins, including enzymes, hormones, hemoglobin, inhibition of cell differentiation, reversion of transformed characteristics of cells to normal morphological and biochemical pattern, increase in interferon antiviral efficiency and induction of integrated viruses. Most if not all these effects of butyrate could result from histone hyperacetylation, from changes in chromatin structures as measured by accessibility to DNases and from modifications in cytoskeleton assembly. We do not know at the present time whether butyrate acts on a very specific target site in cell or if it acts on several cell components.  相似文献   

2.
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   

3.
Wu J  Zhou Z  Hu Y  Dong S 《遗传学报》2012,39(8):375-384
Butyrate has been recently identified as a natural ligand of the G-protein-coupled receptor 41(GPR41).In addition,it is an inhibitor of histone deacetylase(HDAC).Butyrate treatment results in the hyperacetylation of histones,with resultant multiple biological effects including inhibition of proliferation,induction of cell cycle arrest,and apoptosis,in a variety of cultured mammalian cells.However,it is not clear whether GPR41 is actively involved in the above-mentioned processes.In this study,we generated a stable cell line expressing the hGPR41 receptor in order to investigate the involvement of GPR41 on butyrate-induced biochemical and physiologic processes.We found that GPR41 activation may be a compensatory mechanism to counter the increase in histone H3 acetylation levels induced by butyrate treatment.Moreover,GPR41 had an inhibitory effect on the anti-proliferative,pro-apoptotic effects of butyrate.GPR41 expression induced cell cycle arrest at the G1-stage,while its activation by butyrate can cause more cells to pass the G1 checkpoint.These results indicated that GPR41 was associated with histone acetylation and might be involved in the acetylation-related regulation of cell processes including proliferation,apoptosis,and the cell cycle.  相似文献   

4.
The role of histone hyperacetylation in regard to growth, differentiation, and apoptosis in colon cancer cells was assessed in an in vitro model system. HT-29 cells were grown in +/-10% fetal bovine serum with either 5 mM sodium butyrate or 0.3 microM trichostatin A [single dose (T) or 3 doses 8 h apart (TR)] for 24 h. Serum-starved HT-29 cells were further treated with epidermal growth factor or insulin-like growth factor I for an additional 24 h. Apoptosis was quantified with propidium iodide and characterized by electron microscopy. Northern blot analyses were performed with cDNA probes specific for intestinal alkaline phosphatase, Na-K-2Cl cotransporter, the cell cycle inhibitor p21, and the actin control. Flow cytometric analysis revealed a time-dependent growth suppression along with early induction of p21 mRNA in the butyrate, T, and TR groups. Histone hyperacetylation, assessed by acid-urea-triton gel electrophoresis, was transient in the T group but persisted for up to 24 h in the butyrate and TR groups. Induction of apoptosis, growth factor unresponsiveness, and differentiation occurred in the butyrate- and TR-treated cells but not those treated with a single dose of trichostatin A. Thus transient hyperacetylation of histones is sufficient to induce p21 expression and produce cellular growth arrest, but prolonged histone hyperacetylation is required for induction of the programs of differentiation, apoptosis, and growth factor unresponsiveness.  相似文献   

5.
Pretreatment of proliferating D. discoideum amoebae with 10 mM butyrate for at least 8 h (one duplicating time) induced a reversible and dose dependent premature expression of several developmental parameters when the cells were starved in the absence of the fatty acid. The aggregative phase of the morphogenetic cycle was reduced in 2 h and the appearance of mature fruiting bodies and spores took place 4 h earlier as a result of butyrate pretreatment. Some developmentally regulated proteins, such as contact-sites A, cell surface lectins and cyclic AMP phosphodiesterase were also expressed 2 h earlier in butyrate pretreated cells than in controls. The level of extracellular cyclic AMP was reduced in butyrate pretreated cells, while other parameters of cyclic AMP metabolism were not affected. Butyrate also caused a partial inhibition of growth and the hyperacetylation of histone H4 in growing amoeba. These results suggest that butyrate acts as an inducer of differentiation in D. discoideum and can therefore be used as an experimental tool in order to explore regulatory mechanisms operating in slime mold differentiation.Abbreviations MES 2-N-morpholinoethanesulfonate - EDTA ethylendiaminotetracetate - TCA trichloroacetate - DTT dithiothreitol - SDS sodium dodecylsulfate  相似文献   

6.
Sodium butyrate at a concentration of 5mM causes significant hyperacetylation of the core histones in the human breast cancer cell line MCF-7. Histone hyperacetylation was achieved in rapidly-growing cells at 40% confluency after 24 hours in 5mM sodium butyrate. More nearly confluent cells did not reach as high a level of histone hyperacetylation. Upon assaying the estrogen receptors, both cytosolic and KCl-extractable nuclear, we found that butyrate treatment had lowered the estrogen receptor levels in both compartments. To our knowledge this is the first report of an effect of sodium butyrate on estrogen receptor levels.  相似文献   

7.
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   

8.
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   

9.
10.
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   

11.
In thymocytes butyrate and trichostatin A are unable to augment dexamethasone-induced apoptosis. In cultured rat thymocytes the extent of apoptosis induced by dexamethasone alone did not increase by addition of 0.1 - 10 mM butyrate. Even more pronounced was the non-additive interrelationship between dexamethasone and trichostatin A, as trichostatin A-induced apoptosis was not only blocked by the presence of dexamethasone but dexamethasone-induced apoptosis was also partially inhibited in the presence of 0.1 - 0.5 microM trichostatin A. The fact that the non-additive relationship with dexamethasone for apoptosis induction was observed with both histone deacetylase inhibitors suggests that in thymocytes this phenomenon is related to histone acetylation. In contrast to this, in the human T cell-derived leukemia cell line CEM-C7H2, dexamethasone did not block butyrate- or trichostatin A-induced apoptosis; moreover, butyrate, in the concentration range of 0.1 - 1 mM, had a marked synergistic effect on dexamethasone-induced apoptosis. This synergism, however, was not mimicked by trichostatin A, indicating that the effect is not related to histone acetylation but rather due to a pleiotropic effect of butyrate. Furthermore, in CEM-C7H2 cells, at higher concentrations of butyrate (5 - 10 mM) or trichostatin A (0.4 - 0.8 microM), there was a minor but reproducible antagonistic effect of dexamethasone on apoptosis induced by each of the two histone deacetylase inhibitors, suggesting that this antagonistic effect too, is related to histone hyperacetylation.  相似文献   

12.
Histone acetylation of Murine Erythroleukemia Cells (MELC) has been re-examined. It is demonstrated that sodium butyrate causes hyperacetylation of core histones in inducible as well as non-inducible MELC strains. This indicates that histone hyperacetylation per se is not sufficient to activate genes. However, [3H]acetate incorporation into core histones of the inducible MELC line F4N increases after induction of differentiation with dimethylsulfoxide (DMSO), in contrast to the non-inducible variant F4+. Thus histone acetylation may play a role as an auxiliary mechanism for gene activation (and inactivation). In addition, the appearance of a histone H3 variant during differentiation of MELC is reported.  相似文献   

13.
14.
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   

15.
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   

16.
Epigenetic silencing of cancer‐related genes by abnormal methylation and the reversal of this process by DNA methylation inhibitors represents a promising strategy in cancer therapy. As DNA methylation affects gene expression and chromatin structure, we investigated the effects of novel DNMT (DNA methyltransferase) inhibitor, RG108, alone and in its combinations with structurally several HDAC (histone deacetylase) inhibitors [sodium PB (phenyl butyrate) or BML‐210 (N‐(2‐aminophenyl)‐N′phenyloctanol diamine), and all‐trans RA (retinoic acid)] in the human PML (promyelocytic leukaemia) NB4 cells. RG108 at different doses from 20 to 100 μM caused time‐, but not a dose‐dependent inhibition of NB4 cell proliferation without cytotoxicity. Temporal pretreatment with RG108 before RA resulted in a dose‐dependent cell growth inhibition and remarkable acceleration of granulocytic differentiation. Prolonged treatments with RG108 and RA in the presence of HDAC inhibitors significantly increased differentiation. RG108 caused time‐dependent re‐expression of methylation‐silenced E‐cadherin, with increase after temporal or continuous treatments with RG108 and RA, or RA together with PB in parallel, in cell maturation, suggesting the role of E‐cadherin as a possible therapeutic marker. These processes required both PB‐induced hyperacetylation of histone H4 and trimethylation of histone H3 at lysine 4, indicating the cooperative action of histone modifications and DNA methylation/demethylation in derepression of E‐cadherin. This work provides novel experimental evidence of the beneficial role of the DNMT inhibitor RG108 in combinations with RA and HDACIs in the effective differentiation of human PML based on epigenetics.  相似文献   

17.
Butyrate and acetate are bacterial metabolites present in the large intestine lumen. Although butyrate is well known to inhibit the in vitro proliferation of human colon carcinoma cells in a process involving the hyperacetylation of specific nuclear histones, little is known about the possible link between butyrate metabolism and its growth-inhibitory effect. In a previous study (Leschelle et al., 2000, Eur J Biochem 267: 6435-6442), we showed that butyrate accumulates and is metabolized in HT-29 Glc(-/+) cells without increasing oxygen consumption. In the present study, using the same cell line incubated with (14)C-labeled butyrate, we determined that a minor part of (14)C from butyrate was recovered in nuclear histones. Unlike butyrate, acetate exerted no effect on cell growth but was a precursor for overall net histone acetylation. Although butyrate was able to increase the cellular AMP/ADP ratio, it did not affect the ATP cell content or the adenylate charge or the oxidation of endogenous L-glutamine. Butyrate oxidation was found to be markedly sensitive to the presence of other substrates with D-glucose decreasing this oxidation and L-malate stimulating it. Furthermore, in the presence of L-malate, the growth-inhibitory effect of butyrate was significantly weaker than in its absence. From these data, we conclude that the metabolism of butyrate downstream acetyl-CoA synthesis is not involved in the butyrate antiproliferative effect. The suggestion that butyrate metabolism in mitochondria is not used in these cells as a fuel but acts as a regulator of butyrate free concentrations (thus limiting its action upon cellular targets), is discussed.  相似文献   

18.
19.
Histone acetylation is a key modification that regulates chromatin accessibility. Here we show that treatment with butyrate or other histone deacetylase (HDAC) inhibitors does not induce histone hyperacetylation in metaphase-arrested HeLa cells. When compared to similarly treated interphase cells, acetylation levels are significantly decreased in all four core histones and at all individual sites examined. However, the extent of the decrease varies, ranging from only slight reduction at H3K23 and H4K12 to no acetylation at H3K27 and barely detectable acetylation at H4K16. Our results show that the bulk effect is not due to increased or butyrate-insensitive HDAC activity, though these factors may play a role with some individual sites. We conclude that the lack of histone acetylation during mitosis is primarily due to changes in histone acetyltransferases (HATs) or changes in chromatin. The effects of protein phosphatase inhibitors on histone acetylation in cell lysates suggest that the reduced ability of histones to become acetylated in mitotic cells depends on protein phosphorylation.  相似文献   

20.
The effect of sodium butyrate on the nuclear proteins of two Chinese hamster cell lines (V79 and CHO) was studied. Butyrate treatment induces hyperacetylation of core histones in both cell lines, while H1 histone shows a different behavior. In CHO cells H1 is dephosphorylated following butyrate incubation; V79 do not show any change of H1 subtypes. It seems that H1 response to butyrate treatment is cell type dependent. Using silver staining a group of proteins that could be present in vivo in the nucleo-protein complex was also detected.  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