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.
Butyrate is a short chain fatty acid (SCFA) produced by bacterial fermentation of dietary fibers in the colon lumen which severely affects the proliferation of colon cancer cells in in vitro experiments. Although butyrate is able to interfere with numerous cellular targets including cell cycle regulator expression, little is known about butyrate metabolism and its possible involvement in its effect upon colon carcinoma cell growth. In this study, we found that HT-29 Glc-/+ cells strongly accumulated and oxidized sodium butyrate without producing ketone bodies, nor modifying oxygen consumption nor mitochondrial ATP synthesis. HT-29 cells accumulated and oxidized sodium acetate at a higher level than butyrate. However, sodium butyrate, but not sodium acetate, reduced cell growth and increased the expression of the cell cycle effector cyclin D3 and the inhibitor of the G1/S cdk-cyclin complexes p21/WAF1/Cip1, demonstrating that butyrate metabolism downstream of acetyl-CoA synthesis is not required for the growth-restraining effect of this SCFA. Furthermore, HT-29 cells modestly incorporated the 14C-labelled carbon from sodium butyrate into cellular triacylglycerols and phospholipids. This incorporation was greatly increased when D-glucose was present in the incubation medium, corresponding to the capacity of hexose to circulate in the pentose phosphate pathway allowing NADPH synthesis required for lipogenesis. Interestingly, when HT-29 cells were cultured in the presence of sodium butyrate, their capacity to incorporate 14C-labelled sodium butyrate into triacylglycerols and phospholipids was increased more than twofold. In such experimental conditions, HT-29 cells when observed under an electronic microscope, were found to be characterized by an accumulation of lipid droplets in the cytosol. Our data strongly suggest that butyrate acts upon colon carcinoma cells upstream of acetyl-CoA synthesis. In contrast, the metabolism downstream of acetyl-CoA [i.e. oxidation in the tricarboxylic acid (TCA) cycle and lipid synthesis] likely acts as a regulator of butyrate intracellular concentration.  相似文献   

2.
Ammonia decreased metabolism by rat colonic epithelial cells of butyrate and acetate to CO2 and ketones but increased oxidation of glucose and glutamine. Ammonia decreased cellular concentrations of oxaloacetate for all substrates evaluated. The extent to which butyrate carbon was oxidized to CO2 after entering the tricarboxylic acid (TCA) cycle was not significantly influenced by ammonia, suggesting there was no major shift toward efflux of carbon from the TCA cycle. Ammonia reduced entry of butyrate carbon into the TCA cycle, and the proportion of CoA esterified with acetate and butyrate correlated positively with the production of CO2 and ketone bodies. Also, ammonia reduced oxidation of propionate but had no effect on oxidation of 3-hydroxybutyrate. Inclusion of glucose, lactate, or glutamine with butyrate and acetate counteracted the ability of ammonia to decrease their oxidation. In rat colonocytes, it appears that ammonia suppresses short-chain fatty acid (SCFA) oxidation by inhibiting a step before or during their activation. This inhibition is alleviated by glucose and other energy-generating compounds. These results suggest that ammonia may only affect SCFA metabolism in vivo when glucose availability is compromised.  相似文献   

3.
beta-Poly(L-malate) is supposed to function in the storage and transport of histones, DNA polymerases and other nuclear proteins in the giant syncytical cells (plasmodia) of myxomycetes. Here we report on the biosynthesis of [14C]beta-poly(L-malate) from injected L-[14C]malate in the plasmodium of Physarum polycephalum. The effects of KCN, arsenate, adenosine 5'-(alpha, beta-methylene)triphosphate, adenosine 5'-(beta, gamma-methylene)triphosphate, guanosine 5'-(beta, gamma-methylene)triphosphate, desulfo coenzyme A and phenylarsinoxid on beta-poly(L-malate) synthesis were studied after their coinjection with L-[14C]malate. The synthesis was not affected by KCN or desulfo coenzyme A, but was blocked by arsenate and adenosine 5'-(alpha,beta-methylene)triphosphate. The plasmodium lysate catalysed an L-malate-dependent ATP-[32P]pyrophosphate exchange, but was devoid of beta-poly(L-malate) synthetic activity under all experimental conditions tested. The results suggested an extramitochondrial synthesis of beta-poly(L-malate), involving the polymerization of beta-L-malyl-AMP. It is assumed that the lack of synthesis in the lysate is caused by the inactivation of beta-poly(L-malate) polymerase involving a cell injury kinase pathway. Because injected guanosine 5'-(beta, gamma-methylene)triphosphate blocks the synthesis, the injury signal is likely to be GTP dependent.  相似文献   

4.
Isolated rat hepatocytes rapidly utilized [(14)C]palmitate and, in particular, synthesized large amounts of neutral lipids from palmitate. Incorporation into cellular lipids occurred at a linear rate proportional to the medium concentration of fatty acids. Oxidation of [(14)C]palmitate to CO(2) increased with time and was much slower than palmitate esterification. Since [(14)C]acetate and [(14)C]glucose were oxidized to CO(2) at a linear rate, the lag in fatty acid oxidation to CO(2) did not involve enzymatic steps subsequent to acetate formation. The relative contribution of palmitate to esterification and to CO(2) formation depended upon the molar ratio of palmitate to albumin (v) and the length of incubation. Dibutyryl cyclic AMP (1 mM) reduced the oxidation of palmitate and acetate to CO(2) by about 50 and 90%, respectively, but did not alter palmitate esterification. However, equivalent concentrations of sodium butyrate produced similar decreases in CO(2) formation. Dibutyryl cyclic AMP (1 mM) also stimulated palmitate oxidation to water-soluble products, principally ketone bodies, by 50-100%. Sodium butyrate exerted no effect, while monobutyryl cyclic AMP and cyclic AMP both stimulated this pathway significantly. These results indicate that both v and dibutyryl cyclic AMP regulate the metabolism of fatty acids by isolated hepatocytes and suggest that hormonal stimulation of adenyl cyclase controls hepatic lipid metabolism.  相似文献   

5.
In addition to its known effect in suppressing the deacetylation of the nucleosomal core histones, sodium butyrate in the concentration range 0.5 to 15 mM causes a selective inhibition of [32P]phosphate incorporation into histones H1 and H2A of cultured HeLa S3 cells. No commensurate general inhibition of phosphorylation is seen in the non-histone nuclear proteins of butyrate-treated cells, but phosphorylation patterns are altered and 32P-uptake may be stimulated, as well as inhibited, depending upon the protein fraction analyzed. The degree of inhibition of histone phosphorylation in intact cells increases progressively as the butyrate concentration is raised from 0.5 to 15 mM. The effect is time-dependent and fully reversible. Butyrate has no effect on the kinetics of phosphate release from previously phosphorylated histones of cultured cells, nor does it significantly alter the rate of dephosphorylation of 32P-labeled histone H1 by endogenous phosphatases in vitro. Despite the suppression of [32P]phosphate incorporation into histones H1 and H2A of butyrate-treated cells, Na-butyrate does not inhibit the in vitro activities of either type I or type II cyclic AMP-dependent protein kinases, or the cAMP-independent H1 kinase associated with cell cycle progression. This suggests that the butyrate effect on histone phosphorylation in vivo is indirect and may involve an alteration in substrate accessibility or a modulation of systems affecting kinase activities. The poly(ADP)-ribosylation of HeLa histones is not inhibited by 5 mM Na-butyrate. Cells exposed to butyrate show an impaired methylation of lysine and arginine residues in their histones and nuclear hnRNP particles, respectively.  相似文献   

6.
Glucose metabolism and the mechanisms of NADH oxidation by Treponema hyodysenteriae were studied. Under an N2 atmosphere, washed cell suspensions of the spirochete consumed glucose and produced acetate, butyrate, H2, and CO2. Approximately twice as much H2 as CO2 was produced. Determinations of radioactivity in products of [14C]glucose and [14C]pyruvate metabolism and analyses of enzyme activities in cell lysates revealed that glucose was catabolized to pyruvate via the Embden-Meyerhof-Parnas pathway. The results of pyruvate exchange reactions with NaH14CO3 and Na14COOH demonstrated that pyruvate was converted to acetyl coenzyme A (acetyl-CoA), H2, and CO2 by a clostridium-type phosphoroclastic mechanism. NADH:ferredoxin oxidoreductase and hydrogenase activities were present in cell lysates and produced H2 from NADH oxidation. Phosphotransacetylase and acetate kinase catalyzed the formation of acetate from acetyl-CoA. Butyrate was formed from acetyl-CoA via a pathway that involved 3-hydroxybutyryl-coenzyme A (CoA) dehydrogenase, butyryl-CoA dehydrogenase, and butyryl-CoA transferase. T. hyodysenteriae cell suspensions generated less H2 and butyrate under 10% O2-90% N2 than under 100% N2. Cell lysates contained NADH oxidase, NADH peroxidase, and superoxide dismutase activities. These findings indicated there are three major mechanisms that T. hyodysenteriae cells use to recycle NADH generated from the Embden-Meyerhof-Parnas pathway--enzymes in the pathway from acetyl-CoA to butyrate, NADH:ferredoxin oxidoreductase, and NADH oxidase. Versatility in methods of NADH oxidation and an ability to metabolize oxygen could benefit T. hyodysenteriae cells in the colonization of tissues of the swine large bowel.  相似文献   

7.
Glucose metabolism and the mechanisms of NADH oxidation by Treponema hyodysenteriae were studied. Under an N2 atmosphere, washed cell suspensions of the spirochete consumed glucose and produced acetate, butyrate, H2, and CO2. Approximately twice as much H2 as CO2 was produced. Determinations of radioactivity in products of [14C]glucose and [14C]pyruvate metabolism and analyses of enzyme activities in cell lysates revealed that glucose was catabolized to pyruvate via the Embden-Meyerhof-Parnas pathway. The results of pyruvate exchange reactions with NaH14CO3 and Na14COOH demonstrated that pyruvate was converted to acetyl coenzyme A (acetyl-CoA), H2, and CO2 by a clostridium-type phosphoroclastic mechanism. NADH:ferredoxin oxidoreductase and hydrogenase activities were present in cell lysates and produced H2 from NADH oxidation. Phosphotransacetylase and acetate kinase catalyzed the formation of acetate from acetyl-CoA. Butyrate was formed from acetyl-CoA via a pathway that involved 3-hydroxybutyryl-coenzyme A (CoA) dehydrogenase, butyryl-CoA dehydrogenase, and butyryl-CoA transferase. T. hyodysenteriae cell suspensions generated less H2 and butyrate under 10% O2-90% N2 than under 100% N2. Cell lysates contained NADH oxidase, NADH peroxidase, and superoxide dismutase activities. These findings indicated there are three major mechanisms that T. hyodysenteriae cells use to recycle NADH generated from the Embden-Meyerhof-Parnas pathway--enzymes in the pathway from acetyl-CoA to butyrate, NADH:ferredoxin oxidoreductase, and NADH oxidase. Versatility in methods of NADH oxidation and an ability to metabolize oxygen could benefit T. hyodysenteriae cells in the colonization of tissues of the swine large bowel.  相似文献   

8.
Sodium butyrate (SB) treatment was previously shown to produce seven-fold increases in estrogen hormone receptor binding sites of human endometrial adenocarcinoma (IK) cells. Flow cytometric analysis and histone gel electrophoresis were used to examine cell cycle, cell metabolism, and nuclear histone fractions in IK cells treated with different concentrations of SB. SB-treated cells stained with fluorochromes specific for DNA, RNA, or general protein were analyzed by flow cytometry (FCM). Changes in accessibility to three DNA stains and gel electrophoresis were used to analyze rearrangements in chromatin structure. SB caused an accumulation of cells in the G1 phase and inhibited DNA synthesis, but not cellular levels of RNA and protein. Hoechst accessibility to A-T rich regions on DNA was dramatically increased after removal of SB. H1 histones were dephosphorylated and core histones were acetylated during SB-treatment. Information obtained in these studies may be useful for correlating cellular and biochemical events with SB-induced increases in nuclear steroid hormone binding sites.  相似文献   

9.
10.
Summary Active H+ transport in the turtle urinary bladder is mediated by an ATPase. Although the source of ATP is usually mitochondrial oxidative phosphorylation, it is possible because of intracellular compartmentalization or cellular heterogeneity that one metabolic pathway exclusively provides ATP to the pump. To examine this we performed several types of experiments. In one, the coupling between the rate of transport and the rate of oxidation of14C-labeled substrates was studied. We found that there was coupling between H+ transport and glucose, butyrate, oleate, and -OH-butyrate oxidation. In another set of experiments we depleted turtle bladders of their endogenous substrates and tested the effect of a number of substrates on the rate of transport. We found that glucose, pyruvate, lactate, actetate, butyrate and -OH butyrate all stimulated H+ transport. In a third set of experiments we found no coupling between H+ transport and lactate production. Finally, we found that reduction of H+ transport by mucosal acidification resulted in an increase in epithelial cell ATP concentrations and a decrease in ADP levels.These results suggest that the H+ pump receives its ATP from carbohydrate and fatty acid oxidation. The changes in ATP and ADP levels provide an initial explanation for the coupling of H+ transport to the rate of cellular oxidative metabolism.  相似文献   

11.
Cell differentiation is an orderly process that begins with modifications in gene expression. This process is regulated by the acetylation state of histones. Removal of the acetyl groups of histones by specific enzymes (histone deacetylases, HDAC) usually downregulates expression of genes that can cause cells to differentiate, and pharmacological inhibitors of these enzymes have been shown to induce differentiation in several colon cancer cell lines. Butyrate at high (mM) concentration is both a precursor for acetyl-CoA and a known HDAC inhibitor that induces cell differentiation in colon cells. The dual role of butyrate raises the question whether its effects on HT29 cell differentiation are due to butyrate metabolism or to its HDAC inhibitor activity. To distinguish between these two possibilities, we used a tracer-based metabolomics approach to compare the metabolic changes induced by two different types of HDAC inhibitors (butyrate and the non-metabolic agent trichostatin A) and those induced by other acetyl-CoA precursors that do not inhibit HDAC (caprylic and capric acids). [1,2-13C2]-d-glucose was used as a tracer and its redistribution among metabolic intermediates was measured to estimate the contribution of glycolysis, the pentose phosphate pathway and the Krebs cycle to the metabolic profile of HT29 cells under the different treatments. The results demonstrate that both HDAC inhibitors (trichostatin A and butyrate) induce a common metabolic profile that is associated with histone deacetylase inhibition and differentiation of HT29 cells whereas the metabolic effects of acetyl-CoA precursors are different from those of butyrate. The experimental findings support the concept of crosstalk between metabolic and cell signalling events, and provide an experimental approach for the rational design of new combined therapies that exploit the potential synergism between metabolic adaptation and cell differentiation processes through modification of HDAC activity.  相似文献   

12.
In Vitro Lactate Metabolism by Ruminal Ingesta   总被引:1,自引:1,他引:0  
Ruminal ingesta (300 ml) obtained from a fistulated cow fed alfalfa hay (H), 3.6 kg of grain mixture with corn silage fed ad libitum (S), 2.5:1 grain-alfalfa hay mixture (G), or a 2.5:1 grain-alfalfa hay mixture providing 545 g of sodium and calcium lactate daily (L) were incubated for 8 hr with nonpolymerized sodium lactate or 17% polymerized lactic acid neutralized to pH 6.7. Polymerization had no effect on the rate of lactate utilization. The initial rates of lactate metabolism for the H, G, S, and L ingesta were 0.72, 0.95, 1.8, and 3.4 meq per 100 ml of rumen fluid per hr, respectively. Lactate-2-(14)C was incubated for 4 hr with each type of ruminal ingesta. Of the label recovered in the volatile fatty acids (VFA), 74.1, 61.2, 49.3, and 38.9% was recovered in acetate, and 9.4, 19.8, 23.3, and 51.9% was recovered in propionate with H, G, S, and L ingesta, respectively. The balance of label was distributed between butyrate and valerate. The titratable VFA did not follow this pattern of production. With the hay ingesta, lactate metabolism resulted in a net loss of acetate and a large increase in butyrate. Little propionate was produced. The G, S, and L ingesta metabolized lactate to yield progressively more propionate and less butyrate. Evidence was gathered to suggest that acetate was the primary end product of lactate metabolism but that oxidation of lactate to pyruvate dictated the synthesis of butyrate from acetate to maintain an oxidation-reduction balance. It was noted that acetate and butyrate production from lactate was pH-dependent, with acetate production maximal at pH 7.4 and butyrate at 6.2. Propionate production was largely unaffected within this pH range.  相似文献   

13.
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   

14.
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   

15.
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   

16.
Fixation by strain DCB-1 of CO2 carbon into cell material and organic acids occurred during growth on pyruvate both with and without thiosulfate. By using sodium [14C]bicarbonate and sodium [2-14C]pyruvate, the isotopic composition of products and cells was investigated. Up to 70% of cell carbon was derived from CO2. CO2 carbon was also incorporated into succinate, formate, and acetate. Both carbons of acetate underwent exchange reactions with CO2, although the carboxyl-group exchange was twice as fast. Because strain DCB-1 uses CO2 as its major but not sole carbon source while deriving energy from pyruvate metabolism, we describe its metabolism as mixotrophic. Other mixotrophic conditions also supported growth. Lactate or butyrate, which could not support growth in mineral medium, could replace pyruvate as the oxidizable substrate only when acetate was added to the medium.  相似文献   

17.
Fixation by strain DCB-1 of CO2 carbon into cell material and organic acids occurred during growth on pyruvate both with and without thiosulfate. By using sodium [14C]bicarbonate and sodium [2-14C]pyruvate, the isotopic composition of products and cells was investigated. Up to 70% of cell carbon was derived from CO2. CO2 carbon was also incorporated into succinate, formate, and acetate. Both carbons of acetate underwent exchange reactions with CO2, although the carboxyl-group exchange was twice as fast. Because strain DCB-1 uses CO2 as its major but not sole carbon source while deriving energy from pyruvate metabolism, we describe its metabolism as mixotrophic. Other mixotrophic conditions also supported growth. Lactate or butyrate, which could not support growth in mineral medium, could replace pyruvate as the oxidizable substrate only when acetate was added to the medium.  相似文献   

18.
19.
DEAE-Sepharose chromatography of extracts from Zea mays meristematic cells revealed multiple histone acetyltransferase and histone deacetylase enzyme forms. An improved method for nuclear isolation allowed us to discriminate nuclear and cytoplasmic enzymes. Two nuclear histone acetyltransferases, A1 and A2, a cytoplasmic B-enzyme and two nuclear histone deacetylases, HD1 and HD2, have been identified. The histone specificity of the different enzyme forms has been studied in an in vitro system, using chicken erythrocyte histones as substrate. The cytoplasmic histone acetyltransferase B is the predominant enzyme, which acetylates mainly histone H4 and to a lesser extent H2A. The nuclear histone acetyltransferase A1 preferentially acetylates H3 and also H4, whereas enzyme A2 is specific for H3. This substrate specificity was confirmed with homologous Z. mays histones. The two histone deacetylases differ from each other with respect to ionic strength dependence, inhibition by acetate and butyrate, and substrate specificity. The strong inhibitory effect of acetate on histone deacetylases was exploited to distinguish different histone acetyltransferase forms.  相似文献   

20.
Short chain fatty acids (SCFA), principally butyrate, propionate, and acetate, are produced in the gut through the fermentation of dietary fiber by the colonic microbiotica. Butyrate in particular is the preferred energy source for the cells in the colonic mucosa and has been demonstrated to induce apoptosis in colorectal cancer cell lines. We have used proteomics, specifically 2D-DIGE and mass spectrometry, to identify proteins involved in butyrate-induced apoptosis in HCT116 cells and also to identify proteins involved in the development of butyrate insensitivity in its derivative, the HCT116-BR cells. The HCT116-BR cell line was characterized as being less responsive to the apoptotic effects of butyrate in comparison to its parent cell line. Our analysis has revealed that butyrate likely induces a cellular stress response in HCT116 cells characterized by p38 MAPK activation and an endoplasmic reticulum (ER) stress response, resulting in caspase 3/7 activation and cell death. Adaptive cellular responses to stress-induced apoptosis in HCT116-BR cells may be responsible for the development of resistance to apoptosis in this cell line. We also report for the first time additional cellular processes altered by butyrate, such as heme biosynthesis and dysregulated expression of nuclear lamina proteins, which may be involved in the apoptotic response observed in these cell lines.  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