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.
Long-term treatment of NCB-20 cells with sodium butyrate resulted in a marked increase in the specific binding of [3H]D-Ala2,D-Leu5 enkephalin. This increase was concentration and time dependent, with an EC50 of about 480 microM and a maximal effect detected after 3-day treatment. At saturating concentration of butyrate (1 mM) the increase was three- to fourfold of the untreated control. Scatchard analysis revealed that the butyrate effect was due to an increase in the density of the opioid receptor binding sites. Butyrate also induced a smaller (about twofold) increase in the density of muscarinic cholinergic receptor binding assessed by using [3H]quinuclidinyl benzilate, whereas alpha 2-adrenergic receptor binding assessed by using [3H]clonidine was not significantly affected. The butyrate-induced opioid receptor binding could be totally abolished by the presence of cycloheximide, suggesting that the butyrate effect involves synthesis of the receptor protein. Butyrate treatment did not affect basal and prostaglandin E1-stimulated cyclic AMP levels but caused a three- to fourfold decrease in the IC50 of D-Ala2,D-Leu5 enkephalin for attenuating these cyclic AMP levels and approximately 25% increase in the maximal extent of attenuation. In contrast to the butyrate effect, long-term treatment of NCB-20 cells with 1 mM dibutyryl cyclic AMP induced an 80% decrease in the opioid and alpha 2-adrenergic receptor bindings and a 57% loss of muscarinic cholinergic receptor binding. This down-regulation of muscarinic cholinergic receptor binding sites was associated with a 35% decrease of carbachol-induced phosphoinositide breakdown, whereas the receptor up-regulation induced by butyrate was found to increase the carbachol response by about threefold. The differential regulation by butyrate and dibutyryl cyclic AMP suggests that the butyrate effect is mediated by a mechanism independent of intracellular cyclic AMP. The induction by butyrate of opioid-receptors and muscarinic cholinergic receptors in NCB-20 cells may provide a useful system for studying the regulation of gene expression of these receptor proteins.  相似文献   

2.
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   

3.
S Kassis 《Biochemistry》1985,24(20):5666-5672
Exposure of HeLa cells to 5 mM sodium butyrate, but not 0.6 mM, resulted in a more efficient coupling between their beta-adrenergic receptors and the guanine nucleotide binding stimulatory (Ns) component of adenylate cyclase. Both concentrations of the fatty acid, however, caused an increase in receptor number. beta receptors from control and butyrate-treated cells had the same affinity for isoproterenol. Modulation of this affinity by GTP was greatly enhanced, however, in cells treated with 5 mM butyrate compared to untreated and 0.6 mM butyrate treated cells. The concentration of isoproterenol required to half-maximally stimulate adenylate cyclase (Kact) was reduced in cells treated with 5 mM butyrate. In addition, the Kact for GTP in the presence, but not the absence, of isoproterenol was reduced. The effect of butyrate on the coupling between beta receptors and Ns was analyzed in detail by monitoring the activation of Ns by guanine 5'-O-(3-thiotriphosphate) (GTP gamma S) in a two-step assay. In the absence of isoproterenol, Ns from control and 5 mM butyrate treated cells was activated to the same extent with the same time course and Kact for GTP gamma S. In the presence of isoproterenol, Ns from 5 mM butyrate treated cells was activated more rapidly and extensively than Ns from control cells. The Kact for both GTP gamma S and isoproterenol also was reduced. The rate of agonist-mediated activation of Ns was strongly dependent on temperature, which accentuated the differences between 5 mM butyrate treated and control cells. At 4 degrees C, the difference in rate was 8.8-fold.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
Conflicting results have been reported that butyrate in normal piglets leads either to an increase or to a decrease of jejunal villus length, implying a possible effect on the proliferation of enterocytes. No definitive study was found for the biological effects of butyrate in porcine jejunal epithelial cells. The present study used IPEC-J2 cells, a non-transformed jejunal epithelial line to evaluate the direct effects of sodium butyrate on cell proliferation, cell cycle regulation, and apoptosis. Low concentrations (0.5 and 1 mM) of butyrate had no effect on cell proliferation. However, at 5 and 10 mM, sodium butyrate significantly decreased cell viability, accompanied by reduced levels of p-mTOR and PCNA protein. Sodium butyrate, in a dose-dependent manner, induced cell cycle arrest in G0/G1 phase and reduced the numbers of cells in S phase. In addition, relative expression of p21, p27, and pro-apoptosis bak genes, and protein levels of p21Waf1/Cip1, p27Kip1, cyclinD3, CDK4, and Cleave-caspase3 were increased by higher concentrations of sodium butyrate (1, 5, 10 mM), and the levels of cyclinD1 and CDK6 were reduced by 5 and 10 mM butyrate. Butyrate increased the phosphorylated form of the signaling molecule p38 and phosphorylated JNK. In conclusion, the present in vitro study indicated that sodium butyrate inhibited the proliferation of IPEC-J2 cells by inducing cell cycle arrest in the G0/G1 phase of cell cycles and by increasing apoptosis at high concentrations.  相似文献   

5.
The induction of alkaline phosphatase (ALP) by dibutyryl adenosine 3':5'-cyclic monophosphate (Bt2cAMP) was investigated in strain JTC-12 . P3 cells derived from monkey (Maccaca irus) kidney cortex. ALP activity was increased by Bt2cAMP in a dose-dependent manner, reaching a plateau at concentrations higher than 5 mM with the activity being about 4 times that of the controls. The concentration of Bt2cAMP required for half-maximal induction of ALP activity was about 0.8 mM. ALP activity was increased rapidly by Bt2cAMP for the first 5 days and then continued to increase gradually towards a plateau level. Removal of Bt2cAMP from the medium caused a rapid decrease in the activity, suggesting that the induction of ALP activity by Bt2cAMP is reversible. ALP activity was induced synergistically in the presence of 1 mM sodium butyrate together with Bt2cAMP at concentrations from 0.01 to 1 mM. It was also found that in the presence of 1 mM Bt2cAMP, sodium butyrate increased ALP activity in the same manner as Bt2cAMP did in the presence of 1 mM sodium butyrate. Although dexamethasone, a potent glucocorticoid, had no effect on ALP activity in control cells, the hormone suppressed the ALP activity induced by Bt2cAMP in a dose-dependent manner. At concentrations above 0.2 mM, two xanthine derivatives, theophylline and 3-isobutyl-1-methyl-xanthine (IBMX), also inhibited the induction of ALP activity by 1 mM Bt2cAMP. Inhibitors of protein synthesis, cycloheximide (1.5 micrograms/ml) and pactamycin (10 micrograms/ml), as well as inhibitors of RNA synthesis, actinomycin D (2 micrograms/ml) and alpha-amanitin (50 micrograms/ml), suppressed the induction of ALP activity.  相似文献   

6.
A plasmid containing the human preprourokinase gene cDNA under the control of the simian virus 40 early region promoter was introduced into CHO-K1 cells and recombinant cell lines secreting a relatively high level of urokinase were obtained. In the course of studying the effects of various agents on the recombinant cell lines, we found that exposure of recombinant cells to 5 mM butyrate for 24 hours resulted in a 2-3 fold increase in urokinase production. The induction by butyrate was dose-dependent. The half maximal dose was approximately 2 mM; maximal stimulation occurred at 5-10 mM. Cell growth, on the other hand, was inhibited by butyrate concentrations greater than 2.5 mM. The response of cells to butyrate was rapid: a significant increase in urokinase production was observed 6 hours after exposure to 5 mM butyrate. Butyrate treatment increased not only the extracellular level but also the intracellular level of urokinase.  相似文献   

7.
Treatment of primary cultures of adult rat hepatocytes with 5 mM butyrate inhibited the spontaneous decrease in basal activity and mRNA levels of tyrosine aminotransferase (TAT) that occurred during culture (Staecker et al., submitted). We report here that butyrate treatment of primary cultures of rat hepatocytes initially inhibited the induction of TAT. This inhibition was followed by a period of accelerated TAT induction. TAT induction in butyrate-treated primary cultures of adult rat hepatocytes occurred only after metabolism of butyrate by the cultured hepatocytes. The accelerated induction of TAT in hepatocyte cultures treated with sodium butyrate was reflected by increased TAT activity and mRNA levels. Cultured hepatocytes rapidly metabolized butyrate, but the addition of more butyrate into cultures after its initial metabolism resulted in a rapid reduction in TAT activity. These findings indicate that butyrate treatment can affect the expression of TAT in primary hepatocyte cultures in both a positive (increased basal TAT expression) and a negative (inhibition of the induced expression of TAT) manner.  相似文献   

8.
Short-chain fatty acids (SCFAs) accelerate colonic transit. This study examined whether this action was mediated by activation of the peristaltic reflex. SCFAs (acetate, butyrate, or propionate) were applied to the central compartment of a three-compartment flat-sheet preparation of the rat middle to distal colon. The release of serotonin (5-HT), brain-derived neurotrophic factor (BDNF), and CGRP was measured in all three compartments. Ascending contraction and descending relaxation were measured in the orad and caudad compartments. The addition of SCFAs at physiological to supraphysiological concentrations (0.5-100 mM) to the central compartment elicited concentration-dependent ascending contraction and descending relaxation (EC50 approximately 5 mM). At this concentration, SCFAs induced an 8- to 11-fold increase in 5-HT release and a 2- to 3-fold increase in CGRP release in the central compartment only. They had no effect on BDNF release. CGRP release was inhibited by a 5-HT4 but not a 5-HT3 receptor antagonist. Ascending contraction and descending relaxation were also inhibited by 5-HT4 and by CGRP receptor antagonists added to the central compartment. 5-HT and CGRP release, as well as ascending contraction and descending relaxation induced by mechanical stimulation of the mucosa (2-8 strokes), were significantly augmented by 1 mM acetate. Acetate (1 mM) also doubled propulsive velocity in isolated whole segments of the guinea pig colon. In conclusion, chemical stimulation of the mucosa by SCFAs triggers a peristaltic reflex mediated by the release of 5-HT from mucosal cells and activation of 5-HT4 receptors on sensory CGRP-containing nerve terminals. This SCFA-induced peristaltic pathway augments the peristaltic reflex elicited by mechanical stimulation of the mucosa.  相似文献   

9.
The potential role of butyrate to modulate cellular metabolism through integrin receptor led to evaluation of its effect on collagen biosynthesis in cultured fibroblasts. Confluent human dermal fibroblasts were treated with 2 mM and 4 mM of sodium butyrate (NaB) for 48 h. It was found that butyrate induced collagen biosynthesis and prolidase activity independently of α2β1 integrin signaling. The expressions of both α2 and β1integrin subunits as well as integrin-induced activation of focal adhesion kinase (FAK) were not affected in the cells treated with NaB. Since insulin-like growth factor-I (IGF-I) is the most potent stimulator of collagen biosynthesis in fibroblasts, the effect of butyrate on IGF-I receptor (IGF-IR) expression was evaluated. It was found that the exposure of the cells to 4 mM butyrate contributed to a distinct increase in IGF-IR. It was accompanied by a parallel increase in the expression of Sos protein and MAP-kinases (ERK1, ERK2). The data suggests that butyrate-dependent stimulation of collagen biosynthesis in cultured human skin fibroblasts undergoes through IGF-IR signaling.  相似文献   

10.
We hypothesized that voltage-gated K+ (Kv) currents regulate the resting membrane potential (Em), and that serotonin (5-HT) causes Em depolarization by reducing Kv currents in rat mesenteric artery smooth muscle cells (MASMCs). The resting Em was about -40 mV in the nystatin-perforated patch configuration, and the inhibition of Kv currents by 4-aminopyridine caused marked Em depolarization. The inhibition of Ca2+-activated K+ (KCa) currents had no effect on Em. 5-HT (1 microM) depolarized Em by approximately 11 mV and reduced the Kv currents to approximately 63% of the control at -20 mV. Similar 5-HT effects were observed with the conventional whole-cell configuration with a weak Ca2+ buffer in the pipette solution, but not with a strong Ca2+ buffer. In the presence of tetraethylammonium (1mM), 5-HT caused Em depolarization similar to the control condition. These results indicate that the resting Em is largely under the regulation of Kv currents in rat MASMCs, and that 5-HT depolarizes Em by reducing Kv currents in a [Ca2+]i-dependent manner.  相似文献   

11.
As it had been shown that sodium butyrate promoted hyperacetylation of nuclear histones the effect of 5, 10, and 50 mM sodium butyrate on the invitro synthesis of Hb FIc (an acetylated component of Hb F) and Hb F in cord blood reticulocytes was studied. The presence of 5 mM butyrate showed a significant increase in the production of Hb FIc, whereas Hb F synthesis was unaffected. Higher concentrations of butyrate, however, had no stimulatory effect on Hb FIc synthesis. Results of pulse-chase studies indicated that sodium butyrate inhibited the turnover of preformed Hb FIc, probably by inhibiting deacetylase.  相似文献   

12.
13.
The early effects of penconazole (PCZ) at relatively high concentration (10?4 to 5 × 10?4 M) on changes in pH and in titratable acidity of the medium, transmembrane electrical potential difference (Em), electrolyte leakage and cell morphology were investigated in Egeria densa leaves. At the lowest (10?4 M) concentration and in the presence of a very low (10 μM) K+ concentration, triazole induced an early, moderate hyperpolarization of Em, associated with a decrease of net K+ uptake, suggesting some increase in the passive permeability to K+. This Em hyperpolarization was no longer detectable at high (2 mM) K+out concentration. At high PCZ concentrations (3 × 10?4 M and 5 × 10?4 M) the early hyperpolarization detectable in the presence of a low K+out concentration became transient, and was followed by a marked depolarization. PCZ, at these concentrations, suppressed acidification of the medium, stimulated electrolyte leakage and, in the mesophyll cells, induced some shrinking of the cytoplasm and its disconnection from the cell walls. These results are interpreted as due to an early effect of this triazole leading to the disorganization of the plasma membrane.  相似文献   

14.
AIM. This study was designed to examine whether the class I and class IIa histone deacetylase (HDAC) inhibitors, sodium butyrate and sodium valproate alter the expression of human NCOR1 and/or NCOR2 genes coding for N-CoR (nuclear receptor corepressor) and SMRT (silencing mediator for retinoid and thyroid hormone receptors), respectively. METHODS: Human leukemia HL-60 cells were treated for 24 h with 0.5 and 1 mM sodium butyrate, 1 to 3 mM sodium valproate, 1 mcM all-trans retinoic acid (ATRA) or cotreated with 1 mcM ATRA and 0.5 mM sodium butyrate. The acetylation of histones H3 and H4 was analysed by western blotting. The levels of NCOR1 and NCOR2 mRNA were determined by quantitative real-time PCR. Expression of NCF2 gene coding for the NADPH oxidase subunit p67phox was evaluated as a marker of myeloid differentiation. Results. Both butyrate and valproate increased the acetylation of histone H3 at Lys9 and/or Lys14 as well as histone H4 at Lys12. Both HDAC inhibitors caused a significant increase in NCF2 mRNA levels without affecting NCOR1 or NCOR2 mRNA levels. Similarly, ATRA alone or in combination with butyrate induced NCF2 gene expression without any significant influence on the expression of NCOR1 or NCOR2 genes. CONCLUSION: We conclude that inhibitors of class I and class IIa HDACs do not alter the expression of human NCOR1 or NCOR2 genes and that the onset of myeloid differentiation is not accompanied by induction or repression of these genes in HL-60 cells.  相似文献   

15.
16.
The effects of Co2+, Mn2+, and La3+ (2 mM) and verapamil (5 x 10(-6) M) on membrane conductance (Gm) and resting potential (Em) were studied in chick skeletal muscle fibres developing in culture. Cobalt and manganese had no effect on Gm at any time during myogenesis but verapamil caused a decrease in Gm in immature myotubes. This effect diminished with time and was absent by 3 days after myoblast fusion. Lanthanum caused an increase in Gm at all stages of development. All the agents studied caused a significant depolarization of Em. It is concluded that there is no resting calcium conductance in developing skeletal muscle but that there may be a resting sodium conductance which declines with maturation. Lanthanum may increase Gm by displacing membrane-bound calcium and destabilizing membrane structure. All the agents studied were thought to induce depolarization by an inhibitory action on (Na+ + K+)-ATPase.  相似文献   

17.
The effect of estrogen stimulation in vitro on the electrical properties of vascular smooth muscle (VSM), and the concentration of estrogen receptors in VSM were measured in isolated coronary arteries. Microelectrode measurements of the dog coronary artery membrane potential (Em) showed quiescent values of -51 millivolts (mV) and an input resistance (rin) of 10 megohms. Addition by diethylstilbestrol (DES) at 10(-6) M hyperpolarized the membrane to -64 mV and reduced input resistance (rin) to 5 megohms within 15 minutes. Extrapolation of the Em vs. log [K]o curve to zero potential gave similar values of [K]i of around 170 mM in both normal and DES treated muscles suggesting that the DES induced hyperpolarization is not due to increased Na-K pump activity. The 0.5% ethanol vehicle alone had no effect on the membrane potentials. Tetraethylammonium ion (TEA) induced action potentials in the previously quiescent tissue. When DES was applied in the presence of TEA, the membrane potential increased and the action potentials were abolished. Scatchard analysis of the estrogen receptor binding demonstrated both a high and a low affinity receptor for estrogen in the VSM. These data indicate that DES hyperpolarizes the VSM cells by a mechanism other than an increased Na-K pump activity. The mechanism of this increased Em may be due to factors which increase K+ conductance either mediated directly through estrogen interaction with its cytosolic receptors or through some unidentified second mechanism.  相似文献   

18.
Postnatal changes in the enteric nervous system (ENS) are involved in the establishment of colonic motility. In adult rats, butyrate induced neuroplastic changes in the ENS, leading to enhanced colonic motility. Whether butyrate can induce similar changes during the postnatal period remains unknown. Enemas (Na-butyrate) were performed daily in rat pups between postnatal day (PND) 7 and PND 17. Effects of butyrate were evaluated on morphological and histological parameters in the distal colon at PND 21. The neurochemical phenotype of colonic submucosal and myenteric neurons was analyzed using antibodies against Hu, choline acetyltransferase (ChAT), and neuronal nitric oxide synthase (nNOS). Colonic motility and neuromuscular transmission was assessed in vivo and ex vivo. Butyrate (2.5 mM) enemas had no impact on pup growth and histological parameters compared with control. Butyrate did not modify the number of Hu-immunoreactive (IR) neurons per ganglia. A significant increase in the proportion (per Hu-IR neurons) of nNOS-IR myenteric and submucosal neurons and ChAT-IR myenteric neurons was observed in the distal colon after butyrate enemas compared with control. In addition, butyrate induced a significant increase in both nitrergic and cholinergic components of the neuromuscular transmission compared with control. Finally, butyrate increased distal colonic transit time compared with control. We concluded that butyrate enemas induced neuroplastic changes in myenteric and submucosal neurons, leading to changes in gastrointestinal functions. Our results support exploration of butyrate as potential therapy for motility disorders in preterm infants with delayed maturation of the ENS.  相似文献   

19.
Exposure of HeLa cells to sodium butyrate caused an increase in choleragen (cholera toxin) receptors as measured by increased binding of 125I-choleragen to the intact cells. The process was dependent on time and butyrate concentration; maximal increases (over 40-fold) were observed at 48 h and 5 mM sodium butyrate. Other short chain fatty acids were less effective in elevating choleragen receptors in the order: butyrate greater than pentanoate greater than hexanoate greater than propionate. Acetate and isobutyrate had no effect. The increase in toxin receptors caused by butyrate was reversible and occurred in serum-free medium. The affinity of choleragen for control and butyrate-treated HeLa cells appeared to be similar. Butyrate also induced an elevation in choleragen receptors in rat C6 glial and Friend erythroleukemic cells but not in a butyrate-resistant HeLa mutant. The increase observed in Friend cells paralleled the increase in ganglioside GM1 (galactosyl-N-acetylgalactosaminyl-[N-acetylneuraminyl]-galactosylglucosylceramide), the reported choleragen receptor. Although no GM1 could be detected in untreated Hela cells, small amounts were found in cells exposed to butyrate.  相似文献   

20.
Cellular transglutaminase activity was induced in simian virus-transformed human embryonic lung fibroblasts (WI-38 VA13A) by sodium butyrate. The level of enzyme activity approached a maximum by 6 days; 9–11-fold higher in the presence of sodium butyrate (1 mM) than in its absence. The observed increases in cellular transglutaminase activity could be entirely accounted for by equivalent increases in the levels of enzyme protein measured by inhibition enzyme-linked immunosorbent assay. Sodium butyrate also increased the rate of enzyme synthesis, but had no effect on the rate of cellular transglutaminase degradation. The increase in the rate of enzyme synthesis was matched by an increased level of translatable transglutaminase mRNA as measured in a cell-free translation system. Our results suggest that sodium butyrate regulates cellular transglutaminase at the pretranslational level.  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