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.
Bioactive constituents of Artemisia monosperma   总被引:1,自引:0,他引:1  
During a study on the chemistry and biological activity of Kuwaiti plants, new metabolites including 4,6-dihydroxy-3-[3'-methyl-2'-butenyl]-5-[4'-hydroxy-3'-methyl-2'-butenyl]-cinnamic acid (1), the 3R,8R stereoisomer of the C17 polyacetylene dehydrofalcarindiol (2) and a C10 polyacetylene glucoside (3) were characterised by spectroscopic means. Additionally, the previously characterised natural products 1,3R,8R-trihydroxydec-9-en-4,6-yne (4), spathulenol (5) and eriodyctiol-7-methyl ether (6) were also isolated. Compounds 2, 3, and 4 were evaluated for their ability to inhibit the enzyme 12-lipoxygenase and 3 and 4 showed moderate activity at 30 microg/ml. Compound 2 was evaluated against a panel of colorectal and breast cancer cell lines and IC50 values ranged from 5.8 to 37.6 microg/ml. Against a panel of fast-growing mycobacteria and a standard ATCC strain of Staphylococcus aureus, compound 6 exhibited minimum inhibitory concentrations in the range of 64-128 microg/ml.  相似文献   

2.
Electrokinetic chromatography was employed to separate the enantiomers of two novel functionalized ruthenium(II) complexes with different polypyridyl coordination spheres. The use of anionic carboxymethyl-beta-cyclodextrin as chiral mobile phase additive resulted in maximum efficiency and resolution for the enantiomer separation of both transition metal complexes. The syntheses of the [4-(3-hydroxypropyl)-4'-methyl-2,2'-bipyridine]-bis(2,2'-bipyridine)rethenium(II)-bis(tetrafluoroborate) and [4-(3-hydroxypropyl)-4'-methyl-2,2'-bipyridine]-bis(4,4'-dimethyl-2,2'-bypyridine)ruthenium(II)-bis(tetrafluoroborate) complexes and their complete characterization by means of two-dimensional (1)H and (13)C[(1)H] NMR techniques ((1)H-(1)H COSY and (1)H-(13)C HMQC) as well as elemental analyses and MALDI-TOFMS are described in detail. The functionalized complexes can be used as building blocks for further reactions with polymers, biopolymers, surfaces and nanoparticles.  相似文献   

3.
D-(1,5,6-13C3)Glucose (7) has been synthesized by a six-step chemical method. D-(1,2-13C2)Mannose (1) was converted to methyl D-(1,2-13C2)mannopyranosides (2), and 2 was oxidized with Pt-C and O2 to give methyl D-(1,2-13C2)mannopyranuronides (3). After purification by anion-exchange chromatography, 3 was hydrolyzed to give D-(1,2-13C2)mannuronic acid (4), and 4 was converted to D-(5,6-13C2)mannonic acid (5) with NaBH4. Ruff degradation of 5 gave D-(4,5-13C2)arabinose (6), and 6 was converted to D-(1,5,6-13C3)glucose (7) and D-(1,5,6-13C3)mannose (8) by cyanohydrin reduction. D-(2,5,6-13C3)Glucose (9) was prepared from 8 by molybdate-catalyzed epimerization.  相似文献   

4.
A strain of yeast rendered repair deficient by the conditional expression of the RAD52 locus was used to search for natural products capable of damaging DNA. Four ellagic acid derivatives, namely 3,3'-dimethyl-4'-O-beta-D-glucopyranosyl ellagic acid (1), 3,3',4-trimethyl-4'-O-beta-D-glucopyranosyl ellagic acid (2), 3'-methyl-3,4-O,O-methylidene ellagic acid (3) and 3'-methyl-3,4-O,O-methylidene-4'-O-beta-D-glucopyranosyl ellagic acid (4), were identified by this assay as DNA damaging natural principles from several plants, including Alangium javanicum, Anisophyllea apetala, Crypteronia paniculata, Mouririi sp. and Scholtzia parviflora. Although none of the isolated principles mediated frank strand scission of DNA in vitro, all of them potently inhibited the growth of yeast in the absence of expression of RAD52.  相似文献   

5.
Novel zinniol-related compound 3, named porritoxin sulfonic acid, with an isoindoline skeleton was isolated from the culture liquid of Alternaria porri. The structure was determined to be 2-(2"-sulfoethyl)-4-methoxy-5-methyl-6-(3'-methyl-2'-butenyloxy)-2,3-dihydro-1H-isoindol-1-one. The phytotoxic activities of three isoindolines (1-3) were evaluated in a seedling-growth assay against stone leek and lettuce.  相似文献   

6.
This study used in vitro 13C NMR spectroscopy to directly examine bidirectional reactions of the Wood-Werkman cycle involved in central carbon metabolic pathways of dairy propionibacteria during pyruvate catabolism. The flow of [2-13C]pyruvate label was monitored on living cell suspensions of Propionibacterium freudenreichii subsp. shermanii and Propionibacterium acidipropionici under acidic conditions. P. shermanii and P. acidipropionici cells consumed pyruvate at apparent initial rates of 161 and 39 micromol min(-1) g(-1) (cell dry weight), respectively. The bidirectionality of reactions in the first part of the Wood-Werkman cycle was evident from the formation of intermediates such as [3-13C]pyruvate and [3-13C]malate and of products like [2-13C]acetate from [2-13C]pyruvate. For the first time alanine labeled on C2 and C3 and aspartate labeled on C2 and C3 were observed during [2-13C]pyruvate metabolism by propionibacteria. The kinetics of aspartate isotopic enrichment was evidence for its production from oxaloacetate via aspartate aminotransferase. Activities of a partial tricarboxylic acid pathway, acetate synthesis, succinate synthesis, gluconeogenesis, aspartate synthesis, and alanine synthesis pathways were evident from the experimental results.  相似文献   

7.
The biosyntheses of aroma active gamma- and delta-lactones have been previously characterized in yeasts and plants by incubation of labeled fatty acid derivatives. The lactones were considered as end products. Liquid cultures of the lactone-producing yeast Sporobolomyces odorus were used to investigate catabolic pathways of the lactones by incubation of ethyl (+/-)-5-hydroxy(1-(13)C1)decanoate ((13C)-1b) and methyl (+/-)-4-hydroxy(1-(13)C1)decanoate ((13C)-7a). Aliquots of the culture broth were analyzed with GC/MS after CH2N2 derivatization. S. odorus degraded (13C)-1b to 5-oxo(1-(13)C1)decanoic acid ((13C)-2c) and, subsequently, to pentyl (1-(13)C1)pentanedioate ((13C)-3c) and 3-[(1-(13)C1)carboxypropyl] hexanoate ((13C)-4c) by a Baeyer-Villiger-type oxidation (BVO). In addition, the oxidation of (13C)-7a to 4-oxo(1-(13)C1)decanoic acid ((13C)-8c) and a BVO of (13C)-8c to hexyl (1-(13)C1)butanedioate ((13C)-9c) is reported. So far, BVO has been observed in bacteria and some fungi; the data presented indicate a BVO catalyzed by the yeast S. odorus in the course of endogenous lactone metabolism.  相似文献   

8.
Aspergillus fumigatus CY018 was recognized as an endophytic fungus for the first time in the leaf of Cynodon dactylon. By bioassay-guided fractionation, the EtOAc extract of a solid-matrix steady culture of this fungus afforded two new metabolites, named asperfumoid (1) and asperfumin (2), together with six known bioactive compounds including monomethylsulochrin, fumigaclavine C, fumitremorgin C, physcion, helvolic acid and 5alpha,8alpha-epidioxy-ergosta-6,22-diene-3beta-ol as well as other four known compounds ergosta-4,22-diene-3beta-ol, ergosterol, cyclo(Ala-Leu) and cyclo(Ala-Ile). Through detailed spectroscopic analyses including HRESI-MS, homo- and hetero-nuclear correlation NMR experiments (HMQC, COSY, NOESY and HMBC), the structures of asperfumoid and asperfumin were established to be spiro-(3-hydroxyl-2,6-dimethoxyl-2,5-diene-4-cyclohexone-(1,3')-5'-methoxyl-7'-methyl-(1'H, 2'H, 4'H)-quinoline-2',4'-dione) and 5-hydroxyl-2-(6-hydroxyl-2-methoxyl-4-methylbenzoyl)-3,6-dimethoxyl-benzoic methyl ester, respectively. All of the 12 isolates were subjected to in vitro bioactive assays against three human pathogenic fungi Candida albicans, Tricophyton rubrum and Aspergillus niger. As a result, asperfumoid, fumigaclavine C, fumitremorgin C, physcion and helvolic acid were shown to inhibit C. albicans with MICs of 75.0, 31.5, 62.5, 125.0 and 31.5 microg/mL, respectively.  相似文献   

9.
1. The further degradation of a cholic acid (I) metabolite, (4R)-4-[4alpha-(2-carboxyethyl)-3aalpha-hexahydro-7abeta-methyl-5-oxoindan-1beta-yl]valeric acid (IIa), by Corynebacterium equi was investigated. This organism effected ring-opening and gave (4R)-4-[2alpha-(2-carboxyethyl)-3beta-(3-carboxypropionyl)-2beta-methylcyclopent-1beta-yl]valeric acid (VI). The new metabolite was isolated as its trimethyl ester and identified by partical synthesis. It was not utilized by C. equi. 2. (4R)-4[4alpha-(2-Carboxyethyl)-3aalpha-decahydro-8abeta-methyl5-oxa-6-oxoazulen-1beta-yl]valeric acid (IVa), which is a hypothetical initial oxidation product in the above degradation, was not converted by C. equi into the expected metabolite (VI), but into 3 - [2beta - [(2S) - tetrahydro - 5 - oxofur - 2 - yl] - 1beta - methyl - 5 - oxocyclopent - 1alpha - yl]-propionic acid (VIII), the structure of which was established by partial synthesis. 3. Both the possible precursors of the metabolite (VI), an isomer of the epsilon-lactone (IVa), the gamma-lactone (XIa), and the open form of these lactones, the hydroxytricarboxylic acid (V), were also not utilized by C. equi. 4. Under some incubation conditions, C. equi also converted compound (IIa) and 3-(3aalpha-hexahydro-7abeta-methyl-1,5-dioxoindan-4alpha-yl)propionic acid (IIb) into 5-methyl-4-oxo-octane-1,8-dioic acid (III), (4R)-4-(2,3,4,6,6abeta,7,8,9,9aalpha,9bbeta-decahydro-6abeta-methyl-3-oxo-1H-cyclopenta[f]quinolin-7beta-yl)valeric acid (VII) and probably a monohydroxy derivative of compound (IIa) and compound (III), respectively. 5. The possibility that an initial step in the degradation of compound (IIa) by C. equi is oxygenation of the Baeyer-Villiger type, yielding compound (IVa), is discussed. Metabolic pathways of compound (IIa) to compounds (III), (VI), (VII) and (VIII) are also considered.  相似文献   

10.
The pathway of propionate conversion in a syntrophic coculture of Smithella propionica and Methanospirillum hungatei JF1 was investigated by (13)C-NMR spectroscopy. Cocultures produced acetate and butyrate from propionate. [3-(13)C]propionate was converted to [2-(13)C]acetate, with no [1-(13)C]acetate formed. Butyrate from [3-(13)C]propionate was labeled at the C2 and C4 positions in a ratio of about 1:1.5. Double-labeled propionate (2,3-(13)C) yielded not only double-labeled acetate but also single-labeled acetate at the C1 or C2 position. Most butyrate formed from [2,3-(13)C]propionate was also double labeled in either the C1 and C2 atoms or the C3 and C4 atoms in a ratio of about 1:1.5. Smaller amounts of single-labeled butyrate and other combinations were also produced. 1-(13)C-labeled propionate yielded both [1-(13)C]acetate and [2-(13)C]acetate. When (13)C-labeled bicarbonate was present, label was not incorporated into acetate, propionate, or butyrate. In each of the incubations described above, (13)C was never recovered in bicarbonate or methane. These results indicate that S. propionica does not degrade propionate via the methyl-malonyl-coenzyme A (CoA) pathway or any other of the known pathways, such as the acryloyl-CoA pathway or the reductive carboxylation pathway. Our results strongly suggest that propionate is dismutated to acetate and butyrate via a six-carbon intermediate.  相似文献   

11.
A method is described for the preparation of multi-labeled cortisol and cortisone with (13)C and (2)H via the indan synthon method, starting from chiral 11-oxoindanylpropionic acid. [1, 3-(13)C(2)]Acetone was used for the syntheses of [1,2,4, 19-(13)C(4)]cortisol (cortisol-(13)C(4)) and [1,2,4, 19-(13)C(4)]cortisone (cortisone-(13)C(4)), and [1,3-(13)C(2),1,1,1, 3,3,3-(2)H(6)]acetone was for [1,2,4,19-(13)C(4),1,1,19,19, 19-(2)H(5)]cortisol (cortisol-(13)C(4),(2)H(5)) and [1,2,4, 19-(13)C(4),1,1,19,19,19-(2)H(5)]cortisone (cortisone-(13)C(4), (2)H(5)). The chemical shifts for the (13)C and (1)H NMR spectra of cortisol and cortisone were fully assigned.  相似文献   

12.
Chromatographic separation of the CH2Cl2 extract from leaves of Peperomia serpens yielded two chromenes [5-hydroxy-8-(3',7'-dimethylocta-2',6'-dienyl)-2,2,7-trimethyl-2H-1-chromene (1) and 5-hydroxy-8-(3'-methyl-2'-butenyl)-2,2,7-trimethyl-2H-1-chromene-6-carboxylic acid (2)], besides the known chromene [methyl 5-hydroxy-2,2,7-trimethyl-2H-1-chromene-6-carboxylate (3)] and the flavonoid, dihydrooroxylin (4). Their structural elucidation were achieved by spectroscopic analyses. The antifungal activities of the CH2Cl2 extract and the isolated chromenes were measured bioautographically against Cladosporium cladosporioides and C. sphaerospermum, when it was found that the crude extract showed higher activity as compared to the pure compounds.  相似文献   

13.
5-[1'-[3"-Aminoacetyl-2"-methyl-6",8"-dihalosubstitutedquinazolin-4"(3"H)-onyl]-thiosemicarbazido]-2-oxo/thiobarbituric acids 3a-3h and 5-[2'-amino-5'-[3"-aminomethylene-2"-methyl-6",8"-dihalosubstitutedquinazolin-4"(3"H)-onyl]-1',3',4'-thiadiazol-2'-yl]-2-oxo/thiobarbituric acid 5a-5h were prepared by incorporating 1-[3'-aminoacetyl-2'-methyl-6",8"-dihalosubstituted-quinazolin-4'(3'H)-onyl]-thiosemicarbazides 2a-2d and 2-amino-5-[3'-aminomethylene-2'-methyl-6',8'-dihalosubstituted-quinazolin-4'(3'H)-onyl]-1,3,4-thiadiazoles 4a-4 h respectively at 5(th) position of 2-oxo/thiobarbituric acids (via Mannich reaction). All the newly synthesized compounds were screened for their anti-convulsant activity in MES and PTZ models and were compared with standard drugs phenytoin sodium and sodium valproate. Interestingly, these compounds were found to be devoid of sedative and hypnotic activities when tested. Out of the compounds studied, the most active compound 5h, that is 5-[2'-amino-5'-[3"-aminomethylene-2"-methyl-6",8"-dibromoquinazolin-4"(3"H)-onyl]-1',3',4'-thiadiazol-2'-yl]-2-thiobarbituric acid showed activity (90%) more potent than the standard drug.  相似文献   

14.
Wang J  Li W  Li CG  Hu YZ 《AAPS PharmSciTech》2011,12(3):872-878
Photodegradation of fleroxacin is investigated in different injections and solutions. After UV irradiation, fleroxacin was degraded to afford two major products in large-volume injection (specification, 200 mg:100 ml), while degraded to afford another major product in small-volume injection (specification, 200 mg:2 ml). The photodegradation products were detected and isolated by reversed-phase HPLC. Based on the spectral data (FT-IR, MSn, TOF-MS, 1H/13C, DEPT, and 2D NMR), the structures of these products were: 8-fluoro-9-(4-methyl-piperazin-1-yl)-6-oxo-2,3-dihydro-6H-1-oxa-3a-aza-phenalene-5-carboxylic acid (impurity-I); 6-fluoro-1-(2-fluoro-ethyl)-7-(2-methylamino-ethylamino)-4-oxo-1,4-dihydro-quinoline-3-carboxylic acid (impurity-II); and 6,8-difluoro-1-(2-fluoro-ethyl)-7-(2-methylamino-ethylamino)-4-oxo-1,4-dihydro-quinoline-3-carboxylic acid (impurity-III), respectively. Different photodegradation pathways of fleroxacin were proposed, which led to the different stability characteristics of fleroxacin in the injections. The fluorine atom at C8 is more photolabile in dilute injection, so defluorination and cyclization reactions are prone to take place, whereas photo irradiation only cause ring-opening oxidation reaction of piperazine side chain in concentrated injection.  相似文献   

15.
Ex vivo ?(13)C, (2)H? NMR spectroscopy allowed to estimate the relative sizes of neuronal and glial glutamate pools and the relative contributions of (1-(13)C) glucose and (2-(13)C, 2-(2)H(3)) acetate to the neuronal and glial tricarboxylic acid cycles of the adult rat brain. Rats were infused during 60 min in the right jugular vein with solutions containing (2-(13)C, 2-(2)H(3)) acetate and (1-(13)C) glucose or (2-(13)C, 2-(2)H(3)) acetate only. At the end of the infusion the brains were frozen in situ and perchloric acid extracts were prepared and analyzed by high resolution (13)C NMR spectroscopy (90.5 MHz). The relative sizes of the neuronal and glial glutamate pools and the contributions of acetyl-CoA molecules derived from (2-(13)C, (2)H(3)) acetate or (1-(13)C) glucose entering the tricarboxylic acid cycles of both compartments, could be determined by the analysis of (2)H-(13)C multiplets and (2)H induced isotopic shifts observed in the C4 carbon resonances of glutamate and glutamine. During the infusions with (2-(13)C, 2-(2)H(3)) acetate and (1-(13)C) glucose, the glial glutamate pool contributed 9% of total cerebral glutamate being derived from (2-(13)C, 2-(2)H(3)) acetyl-CoA (4%), (2-(13)C) acetyl-CoA (3%) and recycled (2-(13)C, 2-(2)H) acetyl-CoA (2%). The neuronal glutamate pool accounted for 91% of the total cerebral glutamate being mainly originated from (2-(13)C) acetyl-CoA (86%) and (2-(13)C, 2-(2)H) acetyl-CoA (5%). During the infusions of (2-(13)C, 2-(2)H(3)) acetate only, the glial glutamate pool contributed 73% of the cerebral glutamate, being derived from (2-(13)C, 2-(2)H(3)) acetyl-CoA (36%), (2-(13)C, 2-(2)H) acetyl-CoA (27%) and (2-(13)C) acetyl-CoA (10%). The neuronal pool contributed 27% of cerebral glutamate being formed from (2-(13)C) acetyl-CoA (11%) and recycled (2-(13)C, 2-(2)H) acetyl-CoA (16%). These results illustrate the potential of ?(13)C, (2)H? NMR spectroscopy as a novel approach to investigate substrate selection and metabolic compartmentation in the adult mammalian brain.  相似文献   

16.
A chromene and prenylated benzoic acid from Piper aduncum.   总被引:1,自引:0,他引:1  
In addition to nerolidol, 2',6'-dihydroxy-4'-methoxydihydrochalcone, methyl 2,2-dimethyl-8-(3'-methyl-2'-butenyl)-2H-1-chromene-6-carboxylate, methyl 2,2-dimethyl-2H-1-chromene-6-carboxylate and methyl 8-hydroxy-2,2-dimethyl-2H-1-chromene-6-carboxylate, two new natural products were isolated from the leaves of Piper aduncum, 2,2-dimethyl-2H-1-chromene-6-carboxylic acid and 3-(3',7'-dimethyl-2',6'-octadienyl)-4-methoxybenzoic acid. The structures of the isolates were established based on analysis of spectroscopic data, including ES-MS. The DNA-damaging activity of the isolated compounds was also investigated against mutant strains of Saccharomyces cerevisiae.  相似文献   

17.
The unresolved autotrophic CO2 fixation pathways in the sulfur-reducing Archaebacterium Thermoproteus neutrophilus and in the phototrophic Eubacterium Chloroflexus aurantiacus have been investigated. Autotrophically growing cultures were labelled with [1,4-13C1]succinate, and the 13C pattern in cell constituents was determined by 1H- and 13C-NMR spectroscopy of purified amino acids and other cell constituents. In both organisms succinate contributed to less than 10% of cell carbon, the major part of carbon originated from CO2. All cell constituents became 13C-labelled, but different patterns were observed in the two organisms. This proves that two different cyclic CO2 fixation pathways are operating in autotrophic carbon assimilation in both of which succinate is an intermediate. The 13C-labelling pattern in T. neutrophilus is consistent with the operation of a reductive citric acid cycle and rules out any other known autotrophic CO2 fixation pathway. Surprisingly, the proffered [1,4-13C1]succinate was partially converted to double-labelled [3,4-13C2]glutamate, but not to double-labelled aspartate. These findings suggest that the conversion of citrate to 2-oxoglutarate is readily reversible under the growth conditions used, and a reversible citrate cleavage reaction is proposed. The 13C-labelling pattern in C. aurantiacus disagrees with any of the established CO2 fixation pathways; it therefore demands a novel autotrophic CO2 fixation cycle in which 3-hydroxypropionate and succinate are likely intermediates. The bacterium excreted substantial amounts of 3-hydroxypropionate (5 mM) and succinate (0.5 mM) at the end of autotrophic growth. Autotrophically grown Chloroflexus cells contained acetyl-CoA carboxylase and propionyl-CoA carboxylase activity. These enzymes are proposed to be the main CO2-fixing enzymes resulting in malonyl-CoA and methylmalonyl-CoA formation; from these carboxylation products 3-hydroxypropionate and succinate, respectively, can be formed.  相似文献   

18.
The formation of poly(3-hydroxyalkanoates) (PHAs) in Pseudomonas putida KT2442 from various carbon sources was studied by 13C nuclear magnetic resonance spectroscopy, gas chromatography, and gas chromatography-mass spectroscopy. By using [1-13C]decanoate, the relation between beta-oxidation and PHA formation was confirmed. The labeling pattern in PHAs synthesized from [1-13C]acetate corresponded to the formation of PHAs via de novo fatty acid biosynthesis. Studies with specific inhibitors of the fatty acid metabolic pathways demonstrated that beta-oxidation and de novo fatty acid biosynthesis function independently in PHA formation. Analysis of PHAs derived from [1-13C]hexanoate showed that both fatty acid metabolic routes can function simultaneously in the synthesis of PHA. Furthermore, evidence is presented that during growth on medium-chain-length fatty acids, PHA precursors can be generated by elongation of these fatty acids with an acetyl coenzyme A molecule, presumably by a reverse action of 3-ketothiolase.  相似文献   

19.
Placental fatty acid transfer in humans in vivo was studied using stable isotopes. Four pregnant women undergoing cesarean section received 4 h before delivery an oral dose of [(13)C]palmitic acid (PA), [(13)C]oleic acid (OA), [(13)C]linoleic acid (LA), and [(13)C]docosahexaenoic acid (DHA). Maternal blood samples were collected at -4 h (basal), -3 h, -2 h, -1 h, 0 h, and +1 h relative to time of cesarean section. At the time of birth, venous cord blood and placental tissue were collected. Fatty acid composition was determined by gas-liquid chromatography and isotopic enrichment by gas chromatography-combustion-isotope ratio mass spectrometry. (13)C-enrichment of fatty acids in the nonesterified fatty acids (NEFA) of cord plasma tended to be higher than in NEFA of placenta, with statistically significant differences for the nonesterified OA and DHA ([(13)C]PA, 0.024 +/- 0.011 vs. 0.001 +/- 0.001; [(13)C]OA, 0.042 +/- 0.008 vs. 0.005 +/- 0.003; [(13)C]LA, 0.038 +/- 0.010 vs. 0.008 +/- 0.002; [(13)C]DHA, 0.059 +/- 0.009 vs. 0.010 +/- 0.003). The ratio of tracer fatty acid concentrations of placenta to maternal plasma was significantly higher for [(13)C]DHA than for the other fatty acids ([(13)C]PA, 7.1 +/- 1%; [(13)C]OA, 3.8 +/- 0.4%; [(13)C]LA, 9.2 +/- 1.3%; [(13)C]DHA, 25.9 +/- 3.4%). These results suggest that only a part of the placental NEFA participated in fatty acid transfer, and that the placenta showed a preferential accretion of DHA relative to the other fatty acids.  相似文献   

20.
The biochemical pathways involved in acetyl-L-carnitine utilization were investigated in conscious, freely moving rats by 13C NMR spectroscopy. Following 4-h [(1,2-13C2)acetyl]-L-carnitine infusion in fasted animals, the free carnitine levels in serum were increased, and an efflux of unlabelled acetyl-L-carnitine from tissues was observed. [(1,2-13C2)Acetyl]-L-carnitine was found to enter biosynthetic pathways in liver, and the acetyl moiety was incorporated into both cholesterol and 3-hydroxybutyrate carbon skeleton. In accord with the entry of [(1,2-13C2)acetyl]-L-carnitine in the mitochondrial acetylCoA pool associated with tricarboxylic acid cycle, the 13C label was also found in liver glutamate, glutamine, and glutathione. The analysis of the 13C-labelling pattern in 3-hydroxybutyrate and cholesterol carbon skeleton provided evidence that the acetyl-L-carnitine-derived acetylCoA pool used for ketone bodies synthesis in mitochondria was homogeneous, whereas cholesterol was synthesized from two different acetylCoA pools located in the extra- and intramitochondrial compartment, respectively. Furthermore, cholesterol molecules were shown to be preferentially synthesized by the metabolic route involving the direct channelling of CoA-activated mitochondria-derived ketone bodies into 3-hydroxy-3-methylglutarylCoA pathway, prior to equilibration of their acyl groups with extramitochondrial acetylCoA pool via acetoacetylCoA thiolase.  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