首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Synthesis of the specific monosulfates of cholic acid.   总被引:1,自引:0,他引:1  
G Parmentier  H Eyssen 《Steroids》1975,26(6):721-729
The three isomeric cholic acid-monosulfates were synthetized and characterized. Cholic acid-3-sulfate was obtained by reacting cholic acid for 2 min with chlorosulfonic acid in pyridine and chromatography of the resulting bile salt mixture on Sephadex LH-20. The 7- and the 12-monosulfate were prepared by sulfation of the corresponding monohydroxy-diacetates followed by removal of the acetyl groups by alkaline hydrolysis and purification by chromatography on Sephadex LH-20. On TLC in n-butanol-acetic acid-water (10:1:1, v/v) the Rf values were 0.59 for cholic acid-3-sulfate, 0.52 for cholic acid-7-sulfate and 0.48 for cholic acid-12-sulfate. The time required for complete solvolysis at 37 degrees C in acid methanol-acetone (1:9) was 3 h for cholic acid-3-sulfate, 12 h for the 12-monosulfate and 18 h for the 7-monosulfate.  相似文献   

2.
Biliary excretion of tauroursodeoxycholate-3-sulfate in the rat   总被引:1,自引:0,他引:1  
Akimoto K  Sano N  Takikawa H 《Steroids》2001,66(9):701-705
Biliary organic anion excretion is mediated by an ATP-dependent primary active transporter, multidrug resistance protein 2. On the other hand, a multiplicity of canalicular organic anion transport has been suggested. Ursodeoxycholic acid, the 7beta-epimer of chenodeoxycholic acid, is clinically used for various hepatobiliary diseases. In our previous study, the contribution of multidrug resistance protein 2 for biliary excretion of taurine-conjugated bile acid sulfates depended on the numbers of hydroxyl residue. Therefore, to further examine the effect of hydrophobicity on the substrate specificity of multidrug resistance protein 2, we examined the effect of bile acid conjugates and organic anions on biliary excretion of tauroursodeoxycholate-3-sulfate, taurine and sulfonate-conjugated ursodeoxycholic acid, in rats. Biliary tauroursodeoxycholate-3-sulfate excretions was markedly delayed in Eisai hyperbilirubinemic rats. Taurolithocholate-3-sulfate inhibited but ursodeoxycholate-3,7-disulfate did not affect biliary tauroursodeoxycholate-3-sulfate excretion. Biliary tauroursodeoxycholate-3-sulfate excretion was inhibited by sulfobromophthalein, but was not inhibited by dibromosulfophthalein and cefpiramide. These findings indicate that tauroursodeoxycholate-3-sulfate is very specific for multidrug resistance protein 2.  相似文献   

3.
A method is described for the measurement, by difference, of the sulphate fractions of faecal bile acids. A solvolysis step (for the deliberate hydrolysis of the bile acid sulphates) was added to the procedure of sample homogenisation, extraction, enzymatic hydrolysis and thin-layer chromatography. The bile acids were quantitated by gas—liquid chromatography of their methyl ester and trifluoroacetate methyl ester derivatives on 3% QF-1 columns. The total bile acid excretion in 15 control subjects was 603 ± 71 mg/24 h ( ± S.E.M.). The major bile acid peaks (mg/24 h) were: lithocholic acid, without solvolysis 118 ± 26 and including solvolysis 175 ± 30; deoxycholic acid 60 ± 8 and 90 ± 18 and chenodeoxycholic acid 13 ± 7 and 15 ± 7. It was concluded that bile acid sulphates may form a considerable proportion of the total bile acids excreted in man.  相似文献   

4.
When treated with a glucocorticoid, monolayer cultures of a differentiated rat hepatoma convert added chenodeoxycholic and deoxycholic acids into compounds not formed by untreated cultures. The main metabolites have been identified. Chenodeoxycholic acid was 6β-hydroxylated to yield α-muricholic acid, whereas deoxycholic acid was converted mainly into its 3-sulfate. In addition, deoxycholic acid was hydroxylated in 6β-, 7α- and 1ξ-positions. Thus, glucocorticoids appear to induce one 6β-hydroxylase in particular and a sulfotransferase having a notable substrate selectivity.  相似文献   

5.
Ursodeoxycholic acid was estimated in bile samples from humans and wild North American black bears using 7 beta-hydroxysteroid dehydrogenase purified from Clostridium absonum by Procion Red affinity chromatography. The percentage ursodeoxycholic acid was calculated by two methods: (a) 7 beta-hydroxyl groups were quantified using 7 beta-hydroxysteroid dehydrogenase and 3 alpha-hydroxyl groups (total bile acids) were quantified using 3 alpha-hydroxysteroid dehydrogenase. The percentage ursodeoxycholic acid was calculated on the basis of [7 beta-hydroxyl groups]/[3 alpha-hydroxyl groups] X 100. (b) Bile was hydrolyzed with sodium hydroxide and subjected to thin-layer chromatography. Bands corresponding to cholic acid, chenodeoxycholic acid plus deoxycholic acid, and ursodeoxycholic acid were identified by the use of standards and Komarowsky's spray reagent. Total bile acids and total ursodeoxycholic acid were measured by elution of silica gel in unsprayed areas corresponding to the bile acid standards and quantification of the total bile acid in each eluate. Direct comparison of these methods validated the use of 7 beta-hydroxysteroid dehydrogenase in the estimation of ursodeoxycholic acid in the biles of black bears and of patients fed ursodeoxycholic acid for cholesterol gallstone dissolution. Relative percentages of ursodeoxycholic acid were 8-24% in four bears and 22 and 27% in the patients ingesting 500 and 750 mg ursodeoxycholic acid per day for 3 months, respectively. Predictably lower values were obtained in two control subjects and one patient ingesting 750 mg chenodeoxycholic acid per day for 3 months.  相似文献   

6.
Nonsulfated di- to octadeca-saccharides having 2-acetamido-2-deoxy-d-galactose at the reducing end were prepared, in 81% yield, by treatment of chondroitin 6-sulfate (pyridinium salt) with dimethyl sulfoxide containing 10% of water for 14 h at 90°. N-Acetylchondrosine and N-acetyldermosine were obtained from dermatan sulfate of rooster comb, in 30% and 38% yields, respectively, by solvolysis with dimethyl sulfoxide, containing 10% of water, for 30 h at 105°. Hyaluronic acid was also depolymerized by the same solvent in the presence of an equimolar amount of pyridinium sulfate or chloride per disaccharide unit to give reducing di- and higher molecular weight oligo-saccharides. The results of solvolytic desulfation and depolymerization are compared with those of the conventional methods by acid hydrolysis.  相似文献   

7.
Niimura Y  Ishizuka I 《Glycobiology》2006,16(8):729-735
Verots S3 cells derived from the African green monkey kidney were revealed to contain nine types of sulfoglycolipids by incorporating [35S]sulfate. These sulfated glycolipids were separated by DEAE-Sephadex column chromatography and preparative thin-layer chromatography (TLC). The major sulfoglycolipids were characterized using TLC, gas-liquid chromatography (GLC), mass spectrometry, solvolysis, TLC immunostaining, and nuclear magnetic resonance spectra as follows: V1, SM4s (GalCer I3-sulfate); V2, SM3 (LacCer II3-sulfate); V3, SM2a (Gg3Cer II3-sulfate); V4, globopentaosyl ceramide sulfate (Gb5Cer V3-sulfate); V5, (Gg4Cer II3-sulfate, IV3-NeuAc); V6, SB1a (Gg4Cer II3, IV3-bis-sulfate); and V8, (Gg4Cer II3-NeuAc, IV3-sulfate). Both V5 and V8 were sulfated gangliosides comprising both N-acetyl neuraminic acid and sulfate, and this was the first report on V8. A minor component V7 was identified as SM1a (Gg4Cer II3-sulfate) based on its behavior in TLC, GLC, and liquid secondary ion mass spectroscopy. It was postulated that this substance was a precursor of V6 (SB1a) and V5 (Gg4Cer II3-sulfate, IV3-NeuAc), and to date, its presence has not been demonstrated in nature. Another minor component V9 was identified as glucosyl ceramide sulfate based on its migration in TLC and GLC. This renal cell line was shown to be an excellent model for studying the metabolism and function of sulfoglycolipids.  相似文献   

8.
M Numazawa  Y Osawa 《Steroids》1978,32(4):519-527
16alpha-Hydroxyandrostenedione (16alpha-hydroxyandrost-4-ene-3,17-dione), 16alpha-hydroxytestosterone (16alpha,17beta-dihydroxyandrost-4-en-3-one) and 16alpha-hydroxydehydroepiandrosterone 3-sulfate (3beta, 16alpha-dihydroxyandrost-5-en-17-one 3-monosulfate) were synthesized by a new chemical approach with much improved yield. 16alpha-Bromoandrostendione was converted to the hydrazone of 16alpha-hydroxyandrostenedione which gave 16alpha-hydroxyandrostenedione on acid hydrolysis in total 63% yield. Oxidation of 16alpha-hydroxydehydroepiandrosterone with Jones' reagent also selectively afforded 16alpha-hydroxyandrostenedione. 16alpha-Hydroxytestosterone was observed by selective reduction of 16alpha-hydroxyandrostenedione with sodium borohydride. Reaction of 16alpha-hydroxydehydroepiandrosterone with chlorosulfonic acid in pyridine selectively gave the 3-monosulfate. The structure of the sulfate was deduced from its solvolysis to the starting material, and its acetylation and subsequent solvolysis to 16alpha-hydroxydehydroepiandrosterone 16-acetate. All procedures are suitable for large scale synthesis without the use of microorganisms.  相似文献   

9.
The formation of ursodeoxycholic acid from chenodeoxycholic acid and the role of 7-ketolithocholic acid as an intermediate in this biotransformation were studied in vitro in fecal incubations as well as in vivo in the human colon. [24-14C]-Labeled 7-ketolithocholic and chenodeoxycholic acids were studied at various concentrations, and the biotransformation products were analyzed by thin-layer chromatography, gas-liquid chromatography, and mass spectrometry. There was rapid colonic conversion of 7-ketolithocholic acid to ursodeoxycholic acid and, to a lesser extent, to chenodeoxycholic acid. The reduction of 7-ketolithocholic to ursodeoxycholic acid proceeded significantly faster anaerobically and at acid pH than under aerobic and alkaline conditions. When chenodeoxycholic acid was incubated in vitro or instilled into the colon, various amounts of 7-ketolithocholic and ursodeoxycholic acids were formed. The formation of 7-ketolithocholic acid was favored by alkaline conditions. Isotope dilution studies, in which trace amounts of labeled 7-ketolithocholic acid were incubated with unlabeled chenodeoxycholic acid, indicate 7-ketolithocholic acid to be the major intermediate in the intestinal bacterial conversion of chenodeoxycholic to ursodeoxycholic acid.  相似文献   

10.
We developed a highly sensitive and quantitative method to detect bile acid 3-sulfates in human urine employing liquid chromatography/electrospray ionization-tandem mass spectrometry. This method allows simultaneous analysis of bile acid 3-sulfates, including nonamidated, glycine-, and taurine-conjugated bile acids, cholic acid (CA), chenodeoxycholic acid (CDCA), deoxycholic acid (DCA), ursodeoxycholic acid (UDCA), and lithocholic acid (LCA), using selected reaction monitoring (SRM) analysis. The method was applied to analyze bile acid 3-sulfates in human urine from healthy volunteers. The results indicated an unknown compound with the nonamidated common bile acid 3-sulfates on the chromatogram obtained by the selected reaction monitoring analysis. By comparison of the retention behavior and MS/MS spectrum of the unknown peak with the authentic specimen, the unknown compound was identified as 3beta,12alpha-dihydroxy-5beta-cholanoic acid 3-sulfate.  相似文献   

11.
12-Ketochenodeoxycholic acid, an essential intermediate in the synthesis of chenodeoxycholic acid, has been enzymatically prepared from dehydrocholic acid. The specific reduction of dehydrocholic with NADH was catalysed by 3α-hydroxysteroid dehydrogenase (3α-hydroxysteroid: NAD(P)+ oxidoreductase, EC 1.1.1.50) and 7α-hydroxysteroid dehydrogenase (7α-hydroxysteroid:NAD+ 7-oxidoreductase, EC 1.1.1.159). Cofactor regeneration was obtained through the formate dehydrogenase (formate:NAD+ oxidoreductase, EC 1.2.1.2) catalysed oxidation of formate. Complete transformation of dehydrocholic acid to the 12-keto derivative was achieved with a coenzyme turnover number up to 1200. No steroid by-products were detected by high performance liquid chromatography and thin layer chromatography. The process yielded 9 g product l?1 in 66–84 h. The high purity of the enzymatically prepared 12-ketochenodeoxycholic acid should drastically reduce the formation of the toxic by-product lithocholic acid, which occurs in the synthesis of chenodeoxycholic acid when using chemical methods alone.  相似文献   

12.
The gallbladder bile acid composition and the activity of the hepatic steroid 12 alpha-hydroxylase were determined in male and female hamsters. Cholic acid, chenodeoxycholic acid, and deoxycholic acid were the major bile acids in both sexes; in addition, 7-ketodeoxycholic acid and lithocholic acid were present. A sex-linked difference in the ratio of cholic acid (plus its metabolites) to chenodeoxycholic acid (plus its metabolite) was observed. The ratio was 1.93 +/- 0.39 in males and 2.74 +/- 0.54 in females. Another sex-linked difference was found in the activity of the 12 alpha-hydroxylase. The extent of the 12 alpha-hydroxylation of 7 alpha-hydroxycholest-4-en-3-one to yield 7 alpha, 12 alpha-dihydroxycholest-4-en-3-one was about two times greater in the microsomal suspension obtained from the liver of female hamsters than in that of male hamsters. A positive correlation between the 12 alpha-hydroxylase activity and the ratio of cholic acid/chenodeoxycholic acid was also observed. These results strongly support the proposal that the activity of the 12 alpha-hydroxylase is the major factor in determining the relative proportion of cholic acid and chenodeoxycholic acid formed from cholesterol in the liver.  相似文献   

13.
The rate of 7alpha-dehydroxylation of primary bile acids was quantitatively measured radiochromatographically in anaerobically washed whole cell suspensions of Clostridium leptum. The pH optimum for the 7alpha-dehydroxylation of both cholic and chenodeoxycholic acid was 6.5-7.0. Substrate saturation curves were observed for the 7alpha-dehydroxylation of cholic and chenodeoxycholic acid. However, cholic acid whole cell K0.5 (0.37 micron) and V (0.20 mumol hr-1mg protein-1) values differed significantly from chenodeoxycholic acid whole cell K0.5 (0.18 micron) and V (0.50 mumol-1 hr-1 mg protein-1). 7alpha-Dehydroxylation activity was not detected using glycine and taurine-conjugated primary bile acids, ursodeoxycholic acid, cholic acid methyl ester, or hyocholic acid as substrates. Substrate competition experiments showed that cholic acid 7 alpha-dehydroxylation was reduced by increasing concentrations of chendeoxycholic acid; however, chenodeoxycholic acid 7alpha-dehydroxylation activity was unaffected by increasing concentrations of cholic acid. A 10-fold increase in cholic and 7alpha-dehydroxylation activity occurred during the transition from logarithmic to stationary phase growth whether cells were cultured in the presence or absence of sodium cholate. In the same culture, a similar increase in chenodeoxycholic acid 7alpha-dehydroxylation was detected only in cells cultured in the presence of sodium cholate. These results indicate the possible existence of two independent systems for 7alpha-dehydroxylation in C. Leptum.  相似文献   

14.
The formation of chenodeoxycholic acid via 7-ketolithocholic acid by rat liver preparations was examined in vitro. Results showed that a rat liver preparation reduced 7-ketolithocholic acid mainly to chenodeoxycholic acid and to ursodeoxycholic acid in a smaller amount, and that the reductase required NADPH but not NADH as coenzyme and was mainly localized in the microsomes.  相似文献   

15.
7 alpha-Dehydroxylation of cholic acid and chenodeoxycholic acid by whole cells of strain c-25, a Eubacterium lentum-like intestinal anaerobe, was studied. 7 alpha-Dehydroxylase activity was observed only in whole cells grown in the presence of the primary bile acid (cholic acid or chenodeoxycholic acid). Chenodeoxycholic acid was twice as effective as cholic acid as an inducer. Although cells grown in the presence of chenodeoxycholic acid had no significant substrate specificity for the two primary bile acids, cells grown in the presence of cholic acid showed two times greater activity against cholic acid than chenodeoxycholic acid. Exposure of cell suspensions to atmospheric oxygen resulted in little loss of the 7 alpha-dehydroxylase activity. The induced enzyme had an optimal pH range of 7.3 to 7.7. Although adding flavin mononucleotide to the growth medium significantly increased the 7 alpha-dehydroxylation of bile acids without an increase in cell growth, inhibition of the enzyme activity was observed in the resting cell system when flavin mononucleotide was included in the reaction mixture.  相似文献   

16.
7 alpha-Dehydroxylation of cholic acid and chenodeoxycholic acid by whole cells of strain c-25, a Eubacterium lentum-like intestinal anaerobe, was studied. 7 alpha-Dehydroxylase activity was observed only in whole cells grown in the presence of the primary bile acid (cholic acid or chenodeoxycholic acid). Chenodeoxycholic acid was twice as effective as cholic acid as an inducer. Although cells grown in the presence of chenodeoxycholic acid had no significant substrate specificity for the two primary bile acids, cells grown in the presence of cholic acid showed two times greater activity against cholic acid than chenodeoxycholic acid. Exposure of cell suspensions to atmospheric oxygen resulted in little loss of the 7 alpha-dehydroxylase activity. The induced enzyme had an optimal pH range of 7.3 to 7.7. Although adding flavin mononucleotide to the growth medium significantly increased the 7 alpha-dehydroxylation of bile acids without an increase in cell growth, inhibition of the enzyme activity was observed in the resting cell system when flavin mononucleotide was included in the reaction mixture.  相似文献   

17.
1. Using a human hepatoma (Hep G2) cell line that continually synthesizes 3 beta-hydroxy-5-cholenoic acid, lithocholic acid, chenodeoxycholic acid and cholic acid we have determined the metabolism and biological effects of 26-hydroxycholesterol and 7 alpha-hydroxycholesterol. 2. Addition of 26-hydroxycholesterol to the medium (6 microM) downregulated cholesterol and chenodeoxycholic acid synthesis. 3. The predominant metabolite of 26-hydroxycholesterol was 3 beta-hydroxy-5-cholenoic acid. 4. Cholesterol synthesis was not affected by the addition of 7 alpha-hydroxycholesterol (6 and 12 microM). The predominant metabolite of 7 alpha-hydroxycholesterol was chenodeoxycholic acid. 5. In Hep G2 cells 7 alpha-hydroxylation of 26-hydroxycholesterol is not well expressed.  相似文献   

18.
The present report describes the characterization of (24R and 24S)-27-nor-24-methyl-3 alpha, 7 alpha-dihydroxy-5 beta-cholestan-26-oic acids obtained in considerable amounts during the synthesis of (25RS)-3 alpha, 7 alpha-dihydroxy-5 beta-cholestan-26-oic acid by the electrolytic coupling of chenodeoxycholic acid and the half ester of methylsuccinic acid. The mixture of 24R and 24S diastereomers was resolved by analytical and preparative thin-layer chromatography and characterized by gas-liquid chromatography, proton magnetic resonance, and molecular rotation differences. For reference, the model compound, 27-nor-3 alpha, 7 alpha-dihydroxy-5 beta-cholestan-26-oic acid, was synthesized by electrolytic coupling of chenodeoxycholic acid and the half ester of succinic acid.  相似文献   

19.
The isolated livers from normal, streptozotocin-diabetic, and insulin-treated diabetic rats were perfused without and with infused 7 alpha-hydroxycholesterol. Biliary bile acids were extracted and analysed by gas chromatography. In each liver group, total bile acid concentration was more than four times greater with infused 7 alpha-hydroxycholesterol than without the sterol. Without infused 7 alpha-hydroxycholesterol, bile acids in the control group were composed mainly of beta-muricholic acid and to a lesser extent of cholic acid. In the diabetic group, the ratio between these two bile acids reversed. The ratio tended to be normalized by treatment with insulin. With infused 7 alpha-hydroxycholesterol, the control group secreted chenodeoxycholic acid at a considerable higher percentage besides major beta-muricholic acid and minor cholic acid. In the diabetic group, the ratio between the latter two bile acids reversed as was the case with the endogenous secretion, while the percentage of chenodeoxycholic acid remained then unchanged. The diminished percentage of beta-muricholic acid in the diabetic group was increased two times by treatment with insulin.  相似文献   

20.
The hepatic uptake of chenodeoxycholic acid, taurochenodeoxycholic acid, chenodeoxycholic acid 3-sulphate and taurochenodeoxycholate acid 3-sulphate by isolated rat hepatocytes was examined. Taurochenodeoxycholic acid, taurochenodeoxycholic acid 3-sulphate and chenodeoxycholic acid 3-sulphate uptake occurred by a saturable, energy-dependent process while chenodeoxycholic acid uptake was predominantly non-saturable, possibly simple diffusion. Apparent Km (mumol/l) and Vmax (nmol/mg protein per min) values (mean +/- S.D.), respectively, were: chenodeoxycholic acid (saturable component), 33 +/- 6.4 and 4.8 +/- 0.6; taurochenodeoxycholic acid, 11.1 +/- 2.0 and 3.1 +/- 0.5; chenodeoxycholic acid 3-sulphate, 6.1 +/- 0.9 and 2.3 +/- 0.4; and taurochenodeoxycholic acid 3-sulphate, 5.0 +/- 0.7 and 0.9 +/- 0.15. Both conjugation with taurine and sulphation at the 3 position resulted in a reduction in the values of Km and Vmax. Uptake of each of the bile acids taurochenodeoxycholic acid, taurochenodeoxycholic acid 3-sulphate and chenodeoxycholic acid 3-sulphate was competitively inhibited by the other two, with taurochenodeoxycholic acid a potent inhibitor of both taurochenodeoxycholic acid 3-sulphate and chenodeoxycholic acid 3-sulphate uptake. Other bile acids also inhibited. Uptake was inhibited by albumin in the order chenodeoxycholic acid 3-sulphate greater than taurochenodeoxycholic acid 3-sulphate greater than taurochenodeoxycholic acid and was dependent on the extent of bile acid binding to albumin.  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