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.
Tocopherols and tocotrienols are metabolized via hydroxylation and oxidation of their hydrophobic side chain to generate 13′-hydroxychromanols (13′-OHs) and various carboxychromanols, which can be further metabolized by conjugation including sulfation. Recent studies indicate that long-chain carboxychromanols, especially 13′-carboxychromanol (13′-COOH), appear to be more bioactive than tocopherols in anti-inflammatory and anticancer actions. To understand the potential contribution of metabolites to vitamin E-mediated effects, an accurate assay is needed to evaluate bioavailability of these metabolites. Here we describe an LC/MS/MS assay for quantifying vitamin E metabolites using negative polarity ESI. This assay includes a reliable sample extraction procedure with efficacy of ≥ 89% and interday/intraday variation of 3–11% for major metabolites. To ensure accurate quantification, short-chain, long-chain, and sulfated carboxychromanols are included as external/internal standards. Using this assay, we observed that sulfated carboxychromanols are the primary metabolites in the plasma of rodents fed with γ-tocopherol or δ-tocopherol. Although plasma levels of 13′-COOHs and 13′-OHs are low, high concentrations of these compounds are found in feces. Our study demonstrates an LC/MS/MS assay for quantitation of sulfated and unconjugated vitamin E metabolites, and this assay will be useful for evaluating the role of these metabolites in vivo.  相似文献   

2.
Natural forms of vitamin E are metabolized by ω-hydroxylation and β-oxidation of the hydrophobic side chain to generate urinary-excreted 2-(β-carboxyethyl)-6-hydroxychroman (CEHC) and CEHC conjugates (sulfate, glucuronide, or glucoside). We recently showed that sulfated long-chain carboxychromanols, the conjugated intermediate β-oxidation products, are formed from tocopherols and tocotrienols in human cells and in rats. CEHC conjugates have been quantified after being converted to its unconjugated counterpart by sulfatase/glucuronidase. Although the enzymatic hydrolysis is critical for appropriate quantification of conjugated CEHC, it is not clear whether brief incubation of the plasma with sulfatases/glucuronidases results in complete deconjugation of conjugated CEHC. Here we show that quantitative hydrolysis of the conjugated vitamin E metabolites in the plasma requires an extraction procedure using methanol/hexane (2 ml/5 ml) and an overnight sulfatase/glucuronidase hydrolysis. Using this procedure, we demonstrate that conjugated γ-CEHC and some sulfated long-chain carboxychromanols are fully deconjugated. In contrast, direct enzymatic hydrolysis of the whole plasma underestimates the conjugated metabolites by at least threefold. This protocol may be also useful for the analysis of other conjugated phenolic compounds in complicated biological matrices such as plasma.  相似文献   

3.
Although all forms of vitamin E are absorbed, the liver preferentially secretes alpha-, but not gamma-tocopherol, into plasma. Liver alpha-tocopherol secretion is under the control of the alpha-tocopherol transfer protein (TTP). Therefore, to assess gamma-tocopherol bioactivities Ttpa-/-, +/- and +/+ mice were fed for 5 weeks diets containing gamma-tocopherol 550 (gamma-T550), gamma-tocopherol 60 (gamma-T60) mg/kg that also contained trace amounts of alpha-tocopherol, a vitamin E-deficient diet, or a control diet. Plasma and tissues from mice fed gamma-T550 diets were found to contain similar gamma- and alpha-tocopherol concentrations despite the high dietary gamma-tocopherol content; nervous tissues contained almost no gamma-tocopherol. Liver vitamin E metabolites (carboxyethyl hydroxychromans, CEHCs) were also measured. In mice with widely ranging liver alpha- (from 0.7 to 16 nmol/g) and gamma-tocopherol concentrations (0 to 13 nmol/g), hepatic alpha-CEHC was undetectable, but gamma-CEHC concentrations (0.1 to 0.8 nmol/g) were correlated with both alpha- and gamma-tocopherol concentrations (P < 0.004). Hepatic cytochrome P450s (CYPs) involved in vitamin E metabolism, Cyp4f and Cyp3a, were also measured. There were no variations in Cyp4f protein expression as related to diet or mouse genotype. However, Cyp3a was correlated (P < 0.0001) with liver alpha-, but not gamma-tocopherol concentrations. These data support the hypothesis that alpha-tocopherol modulates xenobiotic metabolism by increasing Cyp3a expression, gamma-CEHC formation, and the excretion of both gamma-tocopherol and gamma-CEHC.  相似文献   

4.
Alpha- and gamma-tocopherol (alpha- and gamma-T, respectively) metabolite analysis is of key relevance in the study of vitamin E metabolism. Whilst there is information on urinary excretion of the two major metabolites of these vitamin E homologues, namely the 2,5,7,8-tetramethyl-2-(beta-carboxyethyl)-6-hydroxychroman (alpha-CEHC) and 2,7,8-trimethyl-2-(beta-carboxyethyl)-6-hydroxychroman (gamma-CEHC), their concentration and response to supplements in plasma remains poorly investigated. In this study we describe a gas chromatography-mass spectrometry (GC/MS)-based assay to measure both alpha- and gamma-T and their corresponding CEHC metabolites in human plasma. As an example of the application of this method we report data obtained following the supplemention of two healthy volunteers with 100 mg of deuterium-labeled gamma-T acetate (d(2)-gamma-TAC). Under routine analytical conditions a good linearity in the range 0.0025--1 microM was observed for both the alpha- and gamma-CEHC deuterated standards. In plasma samples, the detection limit for alpha- and gamma-CEHC was 2.5 and 5 nmol/l, respectively. The minimum amount of plasma required for the assay was 500 microl. The plasma concentrations of alpha-CEHC and gamma-CEHC in unsupplemented healthy subjects were 12.6 +/-7.5 and 160.7 +/- 44.9 nmol/l, respectively. In the two volunteers supplemented with 100 mg of d(2)-gamma-TAC, plasma d(2)-gamma-T concentrations increased 250 to 450-fold 6 h postsupplementation. Plasma and urinary d(2)-gamma-CEHC concentrations increased 20 to 40-fold 9--12 h postsupplementation. Interestingly, the acute increase in d(2) gamma-T did not significantly affect the baseline plasma concentrations of d(0)-gamma-T and only slight lowered alpha-T concentrations. Likewise, plasma alpha-CEHC levels were not influenced and urinary excretion of alpha-CEHC were unaltered. This GC/MS method provides a versatile and accurate mean for assessing carboxyethyl-hydroxychroman metabolites of vitamin E in plasma.  相似文献   

5.
There is currently interest in measuring urinary metabolites of vitamin E. It has been suggested that alpha-to-copheronolactone (alphaTL), with an oxidized chroman ring, may be an indicator of in vivo oxidative stress and 2,5,7,8-tetramethyl-2(2'-carboxyethyl)-6-hydroxychroman (alpha-CEHC), with a shortened side chain but intact hydroxychroman ring, may provide a measure of adequate or excess vitamin E status. To date, methods in the literature have tended to concentrate on the estimation of single metabolites. We describe the establishment and validation of a relatively simple and reproducible method to extract and quantitate a range of vitamin E metabolites using 0.5 ml of human urine. The vitamin E metabolites were extracted from urine using solid phase extraction cartridges, deconjugated enzymatically, and analyzed using gas chromatography-mass spectrometry. Using this method we have identified alphaTL and the CEHC metabolites derived from alpha-, delta-, and gamma-tocopherol. In addition we have tentatively identified a novel group of vitamin E metabolites, which are related to the CEHCs but have three extra carbons in the side chain. The possibility of the artifactual oxidation of alpha-CEHC to alphaTL during the assay procedure is also discussed.  相似文献   

6.
Alpha-tocophorel (T) is the most common form of vitamin E inplasma and tissues. Alpha-T is also believed to be superior to its homologues beta-T, gamma-T, and delta-T in antioxidant activity. Biological activity of alpha-T has been intensively studied in a number of bodily systems. In contrast, the other homologues have received little attention beyond the evaluation of their relative antioxidant activity. We as well as others have previously shown that alpha-T can enhance cell- mediated immune function of aged animals and humans. Gamma-T is a principal form of vitamin E in the American diet and some cooking oils contain substantial amount of beta-T and delta-T. Thus it is of public health interest to compare their biological effects with than of alpha-t in various systems. In this study, we used an in vitro supplementation protocol to determine immunologic effects of these T homologues on murine splenocytes. The results showed that all four T homologues enhance both spontaneous and mitogen-stimulated lymphocyte proliferation (LP) and the maximal enhancement produced by them was of the same magnitude. The dose range to produce maximal enhancement varied with different homologues. The efficiency was in the order of beta-T approximately delta-T > alpha-T. Interestingly, at 50 (optimal for alpha-T) and 150 micromol/L, while alpha-T enhanced LP, all the other homologues inhibited LP. This inhibition was found to be due to their cytotoxicity at these levels. T homologues had a differential effect on interleukin (IL)-2 and prostaglandin (PG)E(2) production. IL-2 production by mouse splenocytes was not affected by alpha-T or beta-T, but was increased by gamma-T and delta-T. All T homologues, except for beta-T, inhibited PGE(2) alpha-T. Thus, all the T homologues enhance LP. However, the dose required to reach maximal enhancement varies among the homologues. On the other hand, they have a differential effect on IL-2 and PGE(2) production. The difference in nature and magnitude of the effect on immune function does not correlate with their reported relative antioxidant activity and might be due to minor differences in their structure important to their other biological activities.  相似文献   

7.
The vitamin E family consists of four tocopherols and four tocotrienols. α-Tocopherol (αT) is the predominant form of vitamin E in tissues and its deficiency leads to ataxia in humans. However, results from many clinical studies do not support a protective role of αT in disease prevention in people with adequate nutrient status. On the other hand, recent mechanistic studies indicate that other forms of vitamin E, such as γ-tocopherol (γT), δ-tocopherol, and γ-tocotrienol, have unique antioxidant and anti-inflammatory properties that are superior to those of αT in prevention and therapy against chronic diseases. These vitamin E forms scavenge reactive nitrogen species, inhibit cyclooxygenase- and 5-lipoxygenase-catalyzed eicosanoids, and suppress proinflammatory signaling such as NF-κB and STAT3/6. Unlike αT, other vitamin E forms are significantly metabolized to carboxychromanols via cytochrome P450-initiated side-chain ω-oxidation. Long-chain carboxychromanols, especially 13′-carboxychromanols, are shown to have stronger anti-inflammatory effects than unmetabolized vitamins and may therefore contribute to the beneficial effects of vitamin E forms in vivo. Consistent with mechanistic findings, animal and human studies show that γT and tocotrienols may be useful against inflammation-associated diseases. This review focuses on non-αT forms of vitamin E with respect to their metabolism, anti-inflammatory effects and mechanisms, and in vivo efficacy in preclinical models as well as human clinical intervention studies.  相似文献   

8.
Leukotrienes generated by 5-lipoxygenase (5-LOX)-catalyzed reaction are key regulators of inflammation. In ionophore-stimulated (A23187; 1-2.5 μM) human blood neutrophils or differentiated HL-60 cells, vitamin E forms differentially inhibited leukotriene B(4) (LTB(4)) with an IC(50) of 5-20 μM for γ-tocopherol, δ-tocopherol (δT), and γ-tocotrienol, but a much higher IC(50) for α-tocopherol. 13'-Carboxychromanol, a long-chain metabolite of δT, suppressed neutrophil- and HL-60 cell-generated LTB(4) with an IC(50) of 4-7 μM and potently inhibited human recombinant 5-LOX activity with an IC(50) of 0.5-1 μM. In contrast, vitamin E forms had no effect on human 5-LOX activity but impaired ionophore-induced intracellular calcium increase and calcium influx as well as the subsequent signaling including ERK1/2 phosphorylation and 5-LOX translocation from cytosol to the nucleus, a key event for 5-LOX activation. Further investigation showed that δT suppressed cytosolic Ca(2+) increase and/or LTB(4) formation triggered by ionophores, sphingosine 1-phosphate, and lysophosphatidic acid but not by fMLP or thapsigargin, whereas 13'-carboxychromanol decreased cellular production of LTB(4) regardless of different stimuli, consistent with its strong inhibition of the 5-LOX activity. These observations suggest that δT does not likely affect fMLP receptor-mediated signaling or store depletion-induced calcium entry. Instead, we found that δT prevented ionophore-caused cytoplasmic membrane disruption, which may account for its blocking of calcium influx. These activities by vitamin E forms and long-chain carboxychromanol provide potential molecular bases for the differential anti-inflammatory effects of vitamin E forms in vivo.  相似文献   

9.
Vitamin E has many reported health effects and is recognized as the most important lipid-soluble, chain-breaking antioxidant in the body. Vitamin E has also been reported to play a regulatory role in cell signalling and gene expression. Epidemiological studies show that high blood concentrations of vitamin E are associated with a decreased risk of cardiovascular diseases and certain cancers. Yet, high doses of supplemental vitamin E have been associated with an elevated risk of heart failure and all-cause mortality. Therefore, establishing alternative strategies to improve vitamin E status without potentially increasing mortality risk may prove important for optimal nutrition. To identify dietary phenolic compounds capable of increasing blood and tissue concentrations of vitamin E, selected polyphenols were incorporated into standardized, semi-synthetic diets and fed to male Sprague-Dawley rats for 4 weeks. Blood plasma and liver tissue concentrations of alpha-T and gamma-Twere determined. The flavanols (+)-catechin and (-)-epicatechin, the flavonol quercetin, and the synthetic preservative butylated hydroxytoluene (BHT) markedly elevated the amount of alpha-T in plasma and liver. The sesame lignan sesamin and cereal alkylresorcinols substantially increased the concentrations of gamma-T, but not alpha-T, in the liver. Sesamin also increased gamma-T concentrations in plasma. In order to study the impact of selected polyphenols on the enzymatic degradation of vitamin E, HepG2 cells were incubated together with phenolic compounds in the presence of tocopherols and the formation of metabolites was determined. Sesamin, at concentrations as low as 2 microM, almost completely inhibited tocopherol side-chain degradation and cereal alkylresorcinols inhibited it, dose-dependently (5-20 microM), by 20-80%. BHT, quercetin, (-)-epicatechin, and (+)-catechin had no effect on tocopherol-omega-hydroxylase activity in HepG2 cells. In order to confirm the inhibition of gamma-T metabolism by sesame lignans in humans, sesame oil or corn oil muffins together with deuterium-labelled d6-alpha-Tand d2-gamma-Twere given to volunteers. Urine samples were collected for 72 h and analysed for deuterated and non-deuterated tocopherol metabolites. Consumption of sesame oil muffins significantly reduced the urinary excretion of d2-gamma-CEHC and total (sum of labelled and unlabelled) gamma-CEHC. Overall, the findings from these studies show that the tested dietary phenolic compounds increase vitamin E concentrations through different mechanisms and, thus, have the potential to improve vitamin E status without the use of vitamin E supplements.  相似文献   

10.
Tocopherol vitamers [e.g., alpha-, gamma- and delta-tocopherol (-TOC, γ-TOC and δ-TOC, respectively)] and their water-soluble 2,2′-carboxyethyl hydroxychroman metabolites (e.g., -, γ- and δ-CEHC) all possess antioxidant properties. As a consequence, and similarly to other natural antioxidants, vitamin E compounds may be useful in preventing inflammatory and oxidative-stress-mediated diseases. In this study, we investigated the concentration-dependent effect of tocopherols and water-soluble metabolites on a key event in oxidative stress, for example, the oxidative burst in neutrophils. It was found that not only -TOC but also γ-TOC and δ-TOC as well as -, γ- and δ-CEHC at physiological concentrations inhibit superoxide anion (O2•−) production in phorbol-ester-stimulated neutrophils. This effect was mediated by the inhibition of the translocation and activation of protein kinase C (PKC) enzyme, which is the key event in the phorbol-ester signaling. Importantly, CEHCs were stronger inhibitors of PKC as compared with the vitamer precursors, and the gamma forms of both tocopherol and CEHC showed the highest inhibitory activities. Tocopherols, but not CEHCs, directly inhibit the fully activated nicotine–adenine–dinucleotide phosphate (NADPH) oxidase. However, none of the test compounds was able to directly scavenge O2•− when tested in a cell-free system. In conclusion, vitamin E compounds can control the neutrophil oxidative burst through the negative modulation of PKC-related signaling and NADPH oxidase activity. As an original finding, we observed that CEHC metabolites might contribute to regulate PKC activity in these cells. These results may have important implications in the anti-inflammatory and antioxidant role of vitamin E compounds.  相似文献   

11.
In vitro tocotrienols (T3s) have potent vitamin E antioxidant activity, but unlike tocopherols can inhibit cholesterol synthesis by suppressing 3-hydroxy-3-methyl-glutarylCoA (HMG-CoA) reductase. Because hypercholesterolemia is a major risk factor for coronary artery disease and oxidative modification of low-density lipoprotein (LDL) may be involved in atherogenesis, we investigated whether daily supplements of placebo, or alpha-, gamma-, or delta- (alpha-, gamma-, or delta-) tocotrienyl acetates would alter serum cholesterol or LDL oxidative resistance in hypercholesterolemics in a double-blind placebo controlled study. Subjects were randomly assigned to receive placebo (n = 13), alpha- (n = 13), gamma- (n = 12), or delta- (n = 13) tocotrienyl acetate supplements (250 mg/d). All subjects followed a low-fat diet for 4 weeks, then took supplements with dinner for the following 8 weeks while still continuing diet restrictions. Plasma alpha- and gamma-tocopherols were unchanged by supplementation. Plasma T3s were undetectable initially and always in the placebo group. Following supplementation in the respective groups plasma concentrations were: alpha-T3 0.98 +/- 0.80 micromol/l, gamma-T3 0.54 +/- 0.45 micromol/l, and delta-T3 0.09 +/- 0.07 micromol/l. Alpha-T3 increased in vitro LDL oxidative resistance (+22%, p <.001) and decreased its rate of oxidation (p <. 01). Neither serum or LDL cholesterol nor apolipoprotein B were significantly decreased by tocotrienyl acetate supplements. This study demonstrates that: (i) tocotrienyl acetate supplements are hydrolyzed, absorbed, and detectable in human plasma; (ii) tocotrienyl acetate supplements do not lower cholesterol in hypercholesterolemic subjects on low-fat diets; and (iii) alpha-T3 may be potent in decreasing LDL oxidizability.  相似文献   

12.
The long-chain metabolites of vitamin E (LCM) emerge as a new class of regulatory metabolites and have been considered as the active compounds formed during vitamin E metabolism. The bioactivity of the LCM is comparable to the already established role of other fat-soluble vitamins. The biological modes of action of the LCM are far from being unraveled, but first insights pointed to distinct effects and suggested a specific receptor, which in turn lead to the aforementioned hypothesis. Here, a new facet on the interaction of LCM with foam cell formation of THP-1 macrophages is presented. We found reduced levels of mRNA and protein expression of lipid droplet associated protein PLIN2 by α-tocopherol (α-TOH), whereas the LCM and the saturated fatty acid, stearic acid, increased expression levels of PLIN2. In a lipotoxic setup (0–800?μM stearic acid and 0–100?μM α-TOH or 0–5?μM α-13′-COOH) differences in cellular viability were found. A reduced viability was observed for cells under co-treatment of α-TOH and stearic acid, whereas an increased viability for stearic acid incubation in combination with α-13′-COOH was observed. The striking similarity of PLIN2 expression levels and worsened or mitigated lipotoxicity, respectively, revealed a protective effect of PLIN2 on basal stearic acid-induced lipotoxic conditions in PLIN2 knockdown experiments. Based on our results, we conclude that α-13′-COOH protects cells from lipotoxicity, at least partially via PLIN2 regulation.Herewith another facet of LCM functionality was presented and their reputation as regulatory metabolites was further established.  相似文献   

13.
The Bucherer-Bergs reaction of methyl 2,3-O-isopropylidene-alpha-d-lyxo-hexofuranosid-5-ulose gave (4'S)-4'-carbamoyl-4'-[methyl (4R)-2,3-O-isopropylidene-beta-l-erythrofuranosid-4-C-yl]-oxazolidin-2'-one instead of expected hydantoins. A mixture of hydantoins--(5'R)-triphenylmethoxymethyl-5'-[methyl (4R)-2,3-O-isopropylidene-beta-l-erythrofuranosid-4-C-yl]-imidazolidin-2',4'-dione and (5'S)-triphenylmethoxymethyl-5'-[methyl (4R)-2,3-O-isopropylidene-beta-l-erythrofuranosid-4-C-yl]-imidazolidin-2',4'-dione was obtained from the 5-ulose having protected primary OH group at C-6. The 4'-S configuration of 2 as well as 5'-S configuration of (5'S)-hydroxymethyl-5'-[methyl (4R)-2,3-O-isopropylidene-beta-l-erythrofuranosid-4-C-yl]-imidazolidin-2',4'-dione (9) was confirmed by X-ray crystallography. Corresponding alpha-amino acid--methyl (5S)-5-amino-5-C-carboxy-5-deoxy-alpha-d-lyxo-hexofuranoside (alternative name: 2-[methyl (4R)-beta-l-erythrofuranosid-4-C-yl]-l-serine) (11) was obtained from the hydantoin 9 by acid hydrolysis of the isopropylidene and trityl groups followed by basic hydrolysis of the hydantoin ring. Analogous derivatives with 5-R configuration, formed in a minority, were also isolated and characterised.  相似文献   

14.
从白粉藤(Cissus repens Lank)地上部分分离得到5个木脂素和8个三萜,其中一个木脂素是新化合物,它的结构通过波谱分析和碱水解的方法鉴定为:(+)-异落叶松树脂醇-9′-(2-对-香豆酰)-O-β-D-吡喃木糖苷(1)。其余化合物分别是:(+)-异落叶松树脂醇-9′-O-β-D-吡喃木糖苷(2),(+)-Lyoniside(3),(—)-开环异落叶松树脂醇-9-O-β-D-吡喃木糖苷(4),(7′R,8′S)-4′-hydroxy-3′,5-dimethoxy-7,8′-dihydrobenzofuran-1-propanolneolignan-9′-O-β-D-xylopyranoside(5),木栓酮(6),表木栓醇(7),蒲公英赛醇乙酸酯(8),熊果酸(9),2α-羟基乌索酸(10),积雪草酸(11),Niga-ichigoside F1(12),羽扇豆醇(13)。这些化合物都是首次从该植物中分离得到。  相似文献   

15.
Rats fed with either a sufficient-vitamin A or a vitamin A-free diet were pretreated with 750 mg/kg body weight of retinyl palmitate, alpha-tocopherol acetate, ascorbic acid or glutathione. Benzo[a]pyrene (BaP) metabolism and BaP-induced mutagenesis in Salmonella typhimurium TA98 were investigated and related to lipid peroxidation activities in postmitochondrial (S9) liver fraction. The microsomal mixed-function oxidase activities were decreased by vitamin A deficiency and weakly affected by scavenger treatment. The rate of lipid peroxidation of microsomal membranes was unaffected by vitamin A deficiency because of decreased polyunsaturated fatty acids and increased vitamin E contents. However, lipid peroxidation was decreased by pretreatment with fat-soluble vitamins (chiefly vitamin E) and increased by ascorbic acid. Within each experimental group both BaP metabolism and BaP mutagenic activity were closely correlated with the rate of lipid peroxidation. In vitamin A deficiency, the increased BaP metabolism and mutagenicity could be related to a decrease in cytosolic contents of scavengers (vitamin A and glutathione). In Ames test conditions, the free radical pathway became a route for BaP metabolism and thus the BaP activation to mutagenic metabolites is related to the cellular status in free radical scavengers.  相似文献   

16.
To evaluate vitamin E metabolism, a method was developed to quantitate liver alpha- and gamma-tocopherol metabolites, alpha-carboxyethyl hydroxychroman [alpha-CEHC; 2,5,7,8-tetramethyl-2-(2'-carboxyethyl)-6-hydroxychroman] and gamma-CEHC [2,7,8-trimethyl-2-(2'-carboxyethyl)-6-hydroxychroman], respectively. Vitamin E supraenriched livers were obtained from rats that were injected with vitamin E daily for 18 days. Liver samples (approximately 50 mg) were homogenized, homogenate CEHC-conjugates were hydrolyzed, CEHCs were extracted with ethyl ether, and then CEHCs were quantitated using liquid chromatography-mass spectrometry (LC-MS). Precision, based on intersample variability, ranged from 1% to 3%. Recovery of alpha- and gamma-CEHCs added to liver homogenates ranged from 77% to 87%. Detection limits of alpha- and gamma-CEHC were 20 fmol, with a linear detector response from 0.025 to 20 pmol injected. Corresponding with an increase in liver alpha-tocopherol, the MS peak for liver alpha-CEHC (mass-to-charge ratio 277.8) increased 80-fold (0.18 +/- 0.01 to 15 +/- 2 nmol/g). Liver alpha-CEHC concentrations were correlated with serum alpha-CEHC, liver alpha-tocopherol, and serum alpha-tocopherol (P < 0.001 for each comparison). alpha-CEHC represented 0.5-1% of the liver alpha-tocopherol concentration. Thus, LC-MS can be successfully used to quantitate alpha- and gamma-CEHC in liver samples. These data suggest that in times of excess liver alpha-tocopherol, increased metabolism of alpha-tocopherol to alpha-CEHC occurs.  相似文献   

17.
Tocotrienols (T3s), members of the vitamin E family, exhibit potent anti-cancer, anti-oxidative, anti-inflammatory, and some other biological activities. To better understand the bioavailability and metabolism of T3s, T3s and their metabolites were identified in urine and fecal samples from mice on diet supplemented with mixed T3s using HPLC/electrochemical detection and liquid chromatography electrospray ionisation mass spectrometry (LC–ESI–MS). Whereas the short-chain metabolites carboxyethyl hydroxychromans (CEHCs) and carboxymethylbutyl hydroxychromans (CMBHCs) were the major metabolites of T3s, several new metabolites with double bonds were also identified. Similar to tocopherols, the majority of T3 metabolites were excreted as sulfate/glucuronide conjugates in mouse urine. The distribution of γ- and δ-T3 and γ-T3 metabolites were also determined in different organs as well as in urine and fecal samples from mice on diets supplemented with corresponding T3s. The synergistic anti-cancer actions of γ-T3 and atorvastatin (ATST) were studied in HT29 and HCT116 colon cancer cell lines. The combination greatly potentiated the ability of each individual agent to inhibit cancer cell growth and to induce cell cycle arrest and apoptosis. The triple combination of γ-T3, ATST, and celecoxib exhibited synergistic actions when compared with any double combination plus the third agent. Mechanistic studies revealed that the synergistic actions of γ-T3 and ATST could be attributed to their mediation of 3-hydroxy-3-methyl-glutaryl-CoA reductase, and the subsequent inhibition of protein geranylgeranylation. It remains to be determined whether such a synergy occurs in vivo.  相似文献   

18.
Synthesis of the retinal analog, 10,20-methanoretinal (R6), where the 11Z conformation is locked in a six-membered ring, yielded four stereoisomers (7E,9E,13E, 7E,9E,13Z, 7E,9Z,13E and 7E,9Z,13Z). These four isomers were separated by straight-phase isocratic HPLC and identified by 1H-NMR and NOE analysis. All isomers smoothly recombined with bovine opsin at a relatively high rate (5-10% of that of the natural chromophore 11Z-retinal). The corresponding 13E and 13Z isomers yielded identical analog pigments, probably due to rapid thermal isomerization around the C13 = C14 double bond. The (7E,9E)-isomers produced a pigment with maximal absorbance at 510 nm, while the pigment produced from the (7E,9Z)-isomers had maximal absorbance at 494 nm. Based upon kinetic considerations, the chromophore structure in the 510-nm-absorbing pigment should be (7E,9E,13E), i.e. equivalent to 11Z-retinal and rhodopsin, while the chromophore structure in the 494-nm-absorbing pigment should be (7E,9Z,13Z), i.e. equivalent to (9Z,11Z,13Z)-rhodopsin, an isorhodopsin analog. In analogy to the 11-cis-locked rhodopsin analogs Rh5 and Rh7, the 510-nm-absorbing pigment, (7E,9E,13E)-10,20-methanorhodopsin, was dubbed Rh6 and the 494-nm-absorbing pigment. (7E,9Z,13Z)-10,20-methanorhodopsin, was dubbed Iso6. The opsin shift of Rh6 (2660 cm-1) is practically identical to that of rhodopsin itself (2650 cm-1). Rh6 and Iso6 are nearly as stable as rhodopsin towards hydroxylamine and solubilization in detergent solution and could be easily purified and reconstituted into proteoliposomes by established procedures. Due to the 11-cis-lock, Rh6 is much less photolabile (bleaching rate less than 1%) than rhodopsin, but it is not completely photostable, probably since photoisomerization around the C7 = C8, C9 = C10 and C13 = C14 bonds is allowed. Illumination of either Rh6 or Iso6 does not generate the common photointermediates but instead produces a complex pattern of photochemical transitions, which during continuous illumination leads to the same final steady state, absorbing at 498 nm. This process is accompanied by a slow but steady loss of pigment, probably due to hydrolytic release of chromophore, which is markedly accelerated in the presence of hydroxylamine. In a physiological assay (light-dependent activation of rod cGMP phosphodiesterase) Rh6 is only marginally active and this probably reflects conformational changes accompanying the above-mentioned photochemical transitions. This supports the concept that normal rhodopsin-based phototransduction requires 11Z to all-E isomerization.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

19.
This study aims to investigate in in vivo and in vitro models of nonalcoholic fatty liver disease (NAFLD) the enzymatic metabolism of α-tocopherol (vitamin E) and its relationship to vitamin E-responsive genes with key role in the lipid metabolism and detoxification of the liver. The experimental models included mice fed a high-fat diet combined or not with fructose (HFD+F) and HepG2 human hepatocarcinoma cells treated with the lipogenic agents palmitate, oleate or fructose. CYP4F2 protein, a cytochrome P-450 isoform with proposed α-tocopherol ω-hydroxylase activity, decreased in HFD and even more in HFD+F mice liver; this finding was associated with increased hepatic levels of α-tocopherol and decreased formation of the corresponding long-chain metabolites α-13-hydroxy and α-13-carboxy chromanols. A decreased expression was also observed for PPAR-γ and SREBP-1 proteins, two vitamin E-responsive genes with key role in lipid metabolism and CYP4F2 gene regulation. A transient activation of CYP4F2 gene followed by a repression response was observed in HepG2 cells during the exposure to increasing levels of the lipogenic and cytotoxic agent palmitic acid; such gene repression effect was further exacerbated by the co-treatment with oleic acid and α-tocopherol and was also observed for PPAR-γ and the SREBP isoforms 1 and 2. Such gene response was associated with increased uptake and ω-hydroxylation of α-tocopherol, which suggests a minor role of CYP4F2 in the enzymatic metabolism of vitamin E in HepG2 cells. In conclusion, the liver metabolism and gene response of α-tocopherol are impaired in experimental NAFLD.  相似文献   

20.
3H-Labeled prostaglandins D2 and F2 alpha rapidly degraded to more-polar metabolites in primary cultured rat hepatocytes. The metabolites of prostaglandins D2 and F2 alpha accumulated in the culture medium. The metabolites extracted by ethyl acetate at pH 3 were purified by silicic acid column and thin-layer chromatography of silica gel, and were analysed by gas chromatography-mass spectrometry. The major metabolites from prostaglandin D2 were identified as dinor-prostaglandin D1 (7 alpha,13-dihydroxy-9-ketodinorprost-11-enoic acid) and tetranor-prostaglandin D1 (5 alpha,11- dihydroxy-7-ketotetranorprost-9-enoic acid). Those from prostaglandin F2 alpha were identified as dinor-prostaglandin F1 alpha (7 alpha,9 alpha,13-trihydroxydinorprost-11-enoic acid), tetranor-prostaglandin F1 alpha (5 alpha,7 alpha,11-trihydroxytetranorprost-9-enoic acid) and 9 alpha,11 alpha,15-trihydroxyprost-13-ene-1,20-dioic acid. These data indicate that prostaglandins D2 and F2 alpha mainly degraded by beta-oxidation, which is the same process as reported earlier for prostaglandins E1 and E2, and that prostaglandin F2 alpha was also subjected to omega-oxidation.  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