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.
The positional isomers of the cyclopropane fatty acids of Clostridium butyricum phospholipids have been analyzed by capillary column gas-liquid chromatography. Greater than 95% of the methylenehexadecanoic acids was the 9,10 isomer. On the other hand, 60-70% of the hexadecenoic acid precursors was the Delta(7) isomer, and the remainder was the Delta(9) isomer. Of the methyleneoctadecanoic acids 75-80% was the 11,12 isomer, with the remainder being the 9,10 isomer. There were approximately equal amounts of the Delta(9)- and Delta(11)-octadecenoic acids in the phospholipids. This study reveals a surprisingly strong specificity of the cyclopropane synthetase for the (n-7) series of monoenoic fatty acids. An analysis by capillary column chromatography of the monoenoic and cyclopropane aldehyde dimethylacetals derived from the plasmalogens (1-alk-1'-enyl-2-acyl-glycero-phosphatides) of C. butyricum revealed the presence of the same positional isomeric mixtures of the 16- and 18-carbon monoenoic residues in approximately the same ratios as were found in the fatty acids. In the formation of the cyclopropane alk-1'-enyl ethers there was also specificity for the (n-7) series, but it was not as strong as that seen in the fatty acids. The ratio of the 7,8 isomer to the 9,10 isomer was higher in the methyl-enehexadecanals than in the corresponding fatty acids. This paper extends the use of Golay capillary columns to the analysis of the positional isomers of plasmalogen aldehydes as their dimethylacetal derivatives.  相似文献   

2.
An integrated gas-liquid chromatography method is described for the quantitation of mixtures containing simple monosaccharides in addition to mannuronic, glucuronic, and/or galacturonic acids. A hydrolyzed sample is divided into two portions. One portion is analyzed by the standard aldononitrile method. Glucuronic, galacturonic, and mannuronic acids are converted into compounds that do not chromatograph in the region of the standard aldononitrile acetates. Thus, this analysis gives an accurate estimation of the neutral monosaccharide content. The second portion is analyzed by a modified alditol acetate procedure. The reduction step is repeated three times to convert mannuronic, galacturonic, and glucuronic acids to their corresponding alditols via their intermediate lactones. The results of this gas-liquid chromatography analysis reflect the sum of the monosaccharides present plus their corresponding uronic acids. The difference between the values obtained by the aldononitrile acetate method and the modified alditol acetate method, therefore, is a measure of the uronic acid(s) present.  相似文献   

3.
Thin films of methyl oleate, oleic acid, and di-octadecenoyl phosphatidylcholine were reacted with a 2% solution of OsO4 in water for 1 hr at 0°. As controls, methyl 9,10-dihydroxystearate and 9,10-dihydroxystearic acid were reacted with OsO4 in 0.25 N NaOH in methanol for 1 hr at room temperature. The reaction products were isolated, purified, and analyzed by thin-layer chromatography, gas-liquid chromatography, and infrared and visible spectroscopy. In all cases, the products were identified as diesters of osmic acid in which two molecules of fatty acids are linked through 1 molecule of osmic acid.  相似文献   

4.
Seeds of broad bean (Vicia faba L.) contain a hydroperoxide-dependent fatty acid epoxygenase. Hydrogen peroxide served as an effective oxygen donor in the epoxygenase reaction. Fifteen unsaturated fatty acids were incubated with V. faba epoxygenase in the presence of hydrogen peroxide and the epoxy fatty acids produced were identified. Examination of the substrate specificity of the epoxygenase using a series of monounsaturated fatty acids demonstrated that (Z)-fatty acids were rapidly epoxidized into the corresponding cis-epoxy acids, whereas (E)-fatty acids were converted into their trans-epoxides at a very slow rate. In the series of (Z)-monoenoic acids, the double bond position as well as the chain length influenced the rate of epoxidation. The best substrates were found to be palmitoleic, oleic, and myristoleic acids. Steric analysis showed that most of the epoxy acids produced from monounsaturated fatty acids as well as from linoleic and α-linolenic acids had mainly the (R),(S) configuration. Exceptions were C18 acids having the epoxide group located at C-12/13, in which cases the (S),(R) enantiomers dominated. 13(S)-Hydroxy-9(Z),11(E)-octadecadienoic acid incubated with epoxygenase afforded the epoxy alcohol 9(S),10(R)-epoxy-13(S)-hydroxy-11(E)-octadecenoic acid as the major product. Smaller amounts of the diastereomeric epoxy alcohol 9(R),10(S)-epoxy-13(S)-hydroxy-11(E)-octadecenoic acid as well as the α,β-epoxy alcohol 11(R),12(R)-epoxy-13(S)-hydroxy-9(Z)-octadecenoic acid were also obtained. The soluble fraction of homogenate of V. faba seeds contained an epoxide hydrolase activity that catalyzed the conversion of cis-9,10-epoxyoctadecanoic acid into threo-9,10-dihydroxyoctadecanoic acid.  相似文献   

5.
Seeds of Cichorium intybus L., Crepis thomsonii Babc, and Crepis vesicaria L, were stored from 4 to 8 years at 5°C and then for 18 months under a variety of conditions. Oxygenated acids in Cichorium intybus oil increased from approximately 1% initially to 3% in the first storage period and to 17% while stored at room temperature during the second period. The corresponding levels at these three stages for Crepis thomsonii were 2, 6 and 18%. By gas chromatography (GC) and GC-mass spectrometry, the major oxygenated acids formed during storage were identified as hydroxy acids with conjugated unsaturation and 9,10-epoxy acids. In Crepis vesicaria seed, oil of which contained 53% vernolic (12,13-epoxy-9-octadecenoic) acid originally, approximately 2% of 9,10-epoxides were formed during the storage at room temperature. Levels of hydroxy acids with conjugated unsaturation in this species were 0.3% initially, 2% after 5 years at 5°C, and 9% after 18 months at room temperature. Primary substrates from which oxygenated acids were formed in the three species were crepenynic and linoleic acids, and the almost exclusive formation of 9,10-epoxide from linoleic acid indicated enzymatic involvement.  相似文献   

6.
Allylic hydroxylated derivatives of the C18 unsaturated fatty acids were prepared from linoleic acid (LA) and conjugated linoleic acids (CLAs). The reaction of LA methyl ester with selenium dioxide (SeO2) gave mono-hydroxylated derivatives, 13-hydroxy-9Z,11E-octadecadienoic acid, 13-hydroxy-9E,11E-octadecadienoic acid, 9-hydroxy-10E,12Z-octadecadienoic acid and 9-hydroxy-10E,12E-octadecadienoic acid methyl esters. In contrast, the reaction of CLA methyl ester with SeO2 gave di-hydroxylated derivatives as novel products including, erythro-12,13-dihydroxy-10E-octadecenoic acid, erythro-11,12-dihydroxy-9E-octadecenoic acid, erythro-10,11-dihydroxy-12E-octadecenoic acid and erythro-9,10-dihydroxy-11E-octadecenoic acid methyl esters. These products were purified by normal-phase short column vacuum chromatography followed by high-performance liquid chromatography (HPLC). Their chemical structures were characterized by liquid chromatography-mass spectrometry (LC-MS) and nuclear magnetic resonance spectroscopy (NMR). The allylic hydroxylated derivatives of LA and CLA exhibited moderate in vitro cytotoxicity against a panel of human cancer cell lines including chronic myelogenous leukemia K562, myeloma RPMI8226, hepatocellular carcinoma HepG2 and breast adenocarcinoma MCF-7 cells (IC50 10-75 μM). The allylic hydroxylated derivatives of LA and CLA also showed toxicity to brine shrimp with LD50 values in the range of 2.30-13.8 μM. However these compounds showed insignificant toxicity to honeybee at doses up to 100 μg/bee.  相似文献   

7.
While oat (Avena sativa) has long been known to produce epoxy fatty acids in seeds, synthesized by a peroxygenase pathway, the gene encoding the peroxygenase remains to be determined. Here we report identification of a peroxygenase cDNA AsPXG1 from developing seeds of oat. AsPXG1 is a small protein with 249 amino acids in length and contains conserved heme-binding residues and a calcium-binding motif. When expressed in Pichia pastoris and Escherichia coli, AsPXG1 catalyzes the strictly hydroperoxide-dependent epoxidation of unsaturated fatty acids. It prefers hydroperoxy-trienoic acids over hydroperoxy-dienoic acids as oxygen donors to oxidize a wide range of unsaturated fatty acids with cis double bonds. Oleic acid is the most preferred substrate. The acyl carrier substrate specificity assay showed phospholipid and acyl-CoA were not effective substrate forms for AsPXG1 and it could only use free fatty acid or fatty acid methyl esters as substrates. A second gene, AsLOX2, cloned from oat codes for a 9-lipoxygenase catalyzing the synthesis of 9-hydroperoxy-dienoic and 9-hydroperoxy-trienoic acids, respectively, when linoleic (18:2-9c,12c) and linolenic (18:3-9c,12c,15c) acids were used as substrates. The peroxygenase pathway was reconstituted in vitro using a mixture of AsPXG1 and AsLOX2 extracts from E. coli. Incubation of methyl oleate and linoleic acid or linolenic acid with the enzyme mixture produced methyl 9,10-epoxy stearate. Incubation of linoleic acid alone with a mixture of AsPXG1 and AsLOX2 produced two major epoxy fatty acids, 9,10-epoxy-12-cis-octadecenoic acid and 12,13-epoxy-9-cis-octadecenoic acid, and a minor epoxy fatty acid, probably 12,13-epoxy-9-hydroxy-10-transoctadecenoic acid. AsPXG1 predominately catalyzes intermolecular peroxygenation.  相似文献   

8.
The suberin contents of the isolated superficial cork layers of Malus pumila stems and root ranged from 15 to 35% of the dry weight. The qualitative composition of the aliphatic monomers obtained after alkaline depolymerization of the extractive-free corks was similar but some quantitative differences were found according to cultivar and age of the cork layer. 1-Alkanols (mainly 22:0, 24:0 and 26:0), alkanoic acids (mainly 22:0 and 24:0), α, ω-alkanedioic acids (mainly 16:0, 18:1 (9) and 18:0) and ω-hydroxyalkanoic acids (mainly 18:1 (9) and 22:0) were major constituents of all the samples examined and together they comprised 40–50% of the total monomeric mixture. The remainder was composed mainly of 9,10-epoxy-18-hydroxy-and 9,10,18-trihydroxyoctadecanoic acids. The corresponding dibasic acids, 9,10-epoxy- and 9,10-dihydroxyoctadecane-1,18-dioic, were minor components as were C16 and C18 dihydroxyalkanoic acids (mainly 10,16-dihydroxyhexadecanoic and 10,18-dihydroxyoctadecanoic acids, respectively). The root suberin dittered from that of the stem in containing larger amounts of 9,10-epoxy- 18-hydroxyoctadecanoic and 18-hydroxyoctadec-9-enoic acids.  相似文献   

9.
The syntheses and reactions of two epoxyketoacids (methyl (Z)-9,10-epoxy-13-oxo-(E)-11-octadecenoate (IV) and methyl (E)-9,10-epoxy-13-oxo-(E)-11-octadecenoate (V)) are described. The synthetic method is based on the stereoselective oxidation of linoleic acid by soybean lipoxygenase to produce the corresponding 13-hydroperoxide. Reduction of the hydroperoxide with sodium borohydride followed by oxidation, esterification and epoxidation yielded the compounds IV and V with a global yield of 14% and 3%, respectively, referred to the diasteromerically pure isolated compounds. Confirmation of the structures was carried out by reduction of the ketone group with sodium borohydride and by the opening of the oxirane ring with methanolic boron trifluoride. The reduction of compounds IV and V with hydrogen mainly yielded the tetrahydrofuranoid fatty acid, methyl 10,13-epoxyoctadecanoate. This reaction may be considered a new procedure to obtain tetrahydrofuranoid fatty acids.  相似文献   

10.
Summary From a culture broth ofPseudomonas aeruginosa (KSLA strain 473) grown on heptane as the sole source of carbon, fatty acids could be isolated after a period of decreased oxygen supply. The corresponding methyl esters—obtained by treatment with diazomethane—were separated by gas-liquid chromatography and identified by mass spectrometry. Heptylic, valeric and propionic acids were shown to be present in the original culture broth. Using the same techniques the formation of caproic acid from hexane was shown to occur, whereas the amount of butyric acid formed was extremely small and inconsistent. These results show conclusively that this microbiological oxidation of heptane and hexane proceeds by way of the corresponding fatty acids, which are further degraded by β-oxidation. The absence of caproic and valeric acids in heptane and hexane oxidation, respectively, shows that decarboxylation of fatty acids does not occur.  相似文献   

11.
The effect of bile duct ligation during pregnancy in rats (thereby increasing maternal plasma bile acids levels) on the bile acid content and composition in the fetus was examined. In spite of 30-fold increase in maternal plasma cholic acid, the bile acid content in the fetus of bile duct ligated rats was significantly lower (P <0.05) with a significant reduction in cholic acid content. Plasma cholesterol levels of fetuses from bile duct ligated rats were also significantly lower (p <0.05). In addition to the commonly expected bile acids, gas-liquid Chromatographic analysis of the fetal bile acid pool showed peaks corresponding to several secondary bile acids. These results suggest that the transfer of primary bile acids of maternal origin into the fetus is minimal.  相似文献   

12.
The retinoid-X receptor (RXR) is a ligand activated nuclear receptor that is the heterodimer partner for many class II nuclear receptors. Previously identified natural ligands for this receptor include 9-cis retinoic acid (9cRA), docosahexaenoic acid, and phytanic acid. Our studies were performed to determine if there are any unidentified, physiologically important RXR ligands. Agonists for RXR were purified from rat heart and testes lipid extracts with the use of a cell-based reporter assay to monitor RXR activation. Purified active fractions contained a variety of unsaturated fatty acids and components were quantified by gas-liquid chromatography of derivatized samples. The corresponding fatty acid standards elicited a similar response in the reporter cell assay. Competition binding analysis revealed that the active fatty acids compete with [3H]9cRA for binding to RXR. Non-esterified fatty acids were analyzed from lipid extracts of isolated heart and testes nuclei and endogenous concentrations were found to be within the range of their determined binding affinities. Our studies reveal tissue dependent profiles of RXR agonists and support the idea of unsaturated fatty acids as physiological ligands of RXR.  相似文献   

13.
Five fatty ester derivatives of podophyllotoxin have been prepared by reacting the corresponding fatty acids with the hydroxy group of podophyllotoxin in the presence of dimethylaminopyridine and N,N-dicyclohexylcarbodiimide. The fatty acids incorporated are: 9,12-epoxy-9,11-octadecadienoic acid, octadec-11E-en-9-ynoic acid, 11,12-E-epoxy-octadec-9-ynoic acid, octadeca-9Z,11E-dienoic acid and 9,10-dibromooctadecanoic acid. The average yield of esterification was >95% and the structures of the products were confirmed by a combination of nuclear magnetic resonance spectroscopic and mass spectrometric analyses.  相似文献   

14.
Structural analysis of phosphatidylcholine of plant tissue   总被引:3,自引:0,他引:3  
Pure preparations of phosphatidylcholine were isolated from spinach leaf chloroplasts, spinach leaf microsomes, and cauliflower inflorescence. The isolated phosphatidylcholine was treated with snake venom phospholipase A, and the fatty acid distribution and composition of the fatty acid methyl esters prepared from the lysophosphatidylcholine and the freed fatty acid were determined by gas-liquid chromatography. The results showed that saturated fatty acids were preferentially esterified at position 1 and unsaturated fatty acids at position 2. The phosphatidylcholine from cauliflower was also treated with phospholipase C. The resulting diglycerides were fractionated on AgNO(3)-impregnated thin-layer plates. The diglyceride fractions were transesterified and the fatty acid composition of each was determined by gas-liquid chromatography. The predominant species contained linolenic acid only (22% of the total), linolenic and oleic acids (19%), and linolenic and palmitic acids (37%). These molecular species could not be accounted for by random distribution of the fatty acids.  相似文献   

15.
The distributions of the following monoenoic acids were determined [notation: (position of double bond)-(chain length): (no. of double bonds)]: 7-, 9-, and 11-16:1; 7-, 9-, 11-, and 13-18:1; 9-, 11-, and 13-20:1; 9 + 11-22:1 and 13-22:1. As a rule, all isomers of a group show different distribution patterns. In the phospholipids of fish and mammals, the 7- and 13-isomers of 18:1 accumulate in position 1. In triglycerides of mammals fed on fish they accumulate in positions 1 plus 3, and this distribution is shared by 7-16:1 and 11-16:1 and by the groups 20:1 and 22:1. The positional distribution of the acids seems to depend on their structure, the 9-isomers in general accumulating in position 2; but in triglycerides, at least, the origin of the acid also seems to play a directing role, the exogenous acids being incorporated into positions 1 and 3. The variability of the distribution patterns of 9-16:1, 9-18:1, and 11:18:1, which contrasts with the regularity of the patterns for saturated and polyenoic acids, may be connected with the ability of the endogenous monoenoic acids to balance fluctuations in the supply of the exogenous polyenoic acids, and with the role of the fatty acid 9,10-dehydrogenation mechanism in the maintenance of structural and physical properties of phospholipids and triglycerides.  相似文献   

16.
Four gram-negative bacterial species, including Escherichia coli strain B, Serratia marcescens, Pseudomonas fluorescens, and Vibrio cholerae (comma) strain NIH 41, were investigated for fatty acid content by gas-liquid chromatography involving a preparatory technique which facilitated detection of cyclopropane fatty acids. Methyl esters of fatty acids were subjected to mild catalytic hydrogenation to eliminate unsaturates. Hydrogenation was followed by bromination which removed cyclopropane acids from chromatographic profile patterns. Lactobacillic acid (cis-11,12-methyleneoctanoate) and cis-9,10-methylenehexadecanoate, previously reported in lipids of E. coli and S. marcescens, were found in small amounts in P. fluorescens but were not detected in V. cholerae.  相似文献   

17.
为探究9,10-环甲基十七烷酸(9,10-CMA)诱导灵芝三萜酸合成的相关信号机制,分析了NO信号的介导作用。结果表明,在NO信号分子介导下,9,10-CMA可有效地刺激灵芝菌丝体中三萜合成关键酶细胞色素CYP450和苯丙氨酸解氨酶(PAL)的活化以及三萜酸的合成。NO可作为灵芝菌丝体中CYP450产生和PAL活化的上游分子发挥作用。在9,10-CMA诱导下,通过荧光定量PCR分析了6个关键酶基因在灵芝三萜酸合成过程的动态表达。结果表明,在9,10-CMA诱导下与灵芝三萜酸合成相关的角鲨烯合成酶基因sqs、细胞色素P450单加氧酶CYP5150L8基因和3‐羟基‐3‐甲基戊二酰辅酶A还原酶基因hmgr的上调最显著,提示这3个酶在三萜诱导过程中具有重要作用。  相似文献   

18.
Three O-acylated, unsaturated sialic acids, N-acetyl-9-O-acetyl-, N-acetyl-9-O-lactoyl-, and 2-deoxy-N-glycoloyl-9-O-lactoyl-2,3-didehydroneuraminic acid (5-acetamido-9-O-acetyl-, 5-acetamido-9-O-lactoyl-, and 2,6-anhydro-3,5-dideoxy-5-glycoloylamido-9-O-lactoyl-D-glycero-D-g alacto-non-2- enonic acid) were isolated from urine or submandibular glands of rat, pig, and cow. Mass spectrometric evidence for the existence of 2,3-unsaturated 9-O-acetyl-N-glycoloylneuraminic acid in porcine urine was also obtained. The sialic acids were purified by dialysis, gel- and ion-exchange chromatography, and preparative thin-layer chromatography. They were analyzed by thin-layer chromatography, high-pressure liquid chromatography, and capillary gas-liquid chromatography-mass spectrometry. For comparison, O-acetylated unsaturated sialic acids were synthesized.  相似文献   

19.
Incubation of linoleic acid with the 105,000g particle fraction of the homogenate of the broad bean (Vicia faba L.) led to the formation of the following products: 13(S)-hydroxy-9(Z),11(E)-octadecadienoic acid, 9,10-epoxy-12(Z)-octadecenoic acid (9(R),10(S)/9(S)/10(R), 80/20), 12,13-epoxy-9(Z)-octadecenoic acid (12(S),13(R)/12(R)/13(S), 64/36), and 9,10-epoxy-13(S)-hydroxy-11(E)-octadecenoic acid (9(S),10(R)/9(R),10(S), 91/9). Oleic acid incubated with the enzyme preparation in the presence of 13(S)-hydroperoxy-9(Z),11(E)-octadecadienoic acid or cumene hydroperoxide was converted into 9,10-epoxyoctadecanoic acid (9(R),10(S)/9(S),10(R), 79/21). Two enzyme activities were involved in the formation of the products, an omega 6-lipoxygenase and a hydroperoxide-dependent epoxygenase. The lipoxygenase, but not the epoxygenase, was inhibited by low concentrations of 5,8,11,14-eicosatetraynoic acid and nordihydroguaiaretic acid. In contrast, the epoxygenase, but not the lipoxygenase, was readily inactivated in the presence of 13(S)-hydroperoxy-9(Z),11(E)-octadecadienoic acid. Studies with 18O2-labeled 13(S)-hydroperoxy-9(Z),11(E)-octadecadienoic acid showed that the epoxide oxygens of 9,10-epoxyoctadecanoic acid and of 9,10-epoxy-13(S)-hydroxy-11(E)-octadecenoic acid were derived from hydroperoxide and not from molecular oxygen.  相似文献   

20.
Products of linoleic hydroperoxide-decomposing enzyme of alfalfa seed   总被引:2,自引:0,他引:2  
Alfalfa seeds and seedlings contain an enzyme that catalyzes a reaction with the 13- and 9-hydroperoxides of linoleic acid to form 13-hydroxy-10-oxo-trans-octadecenoic acid and 9-hydroxy-12-oxo-trans-10-octadecenoic acid, respectively. When commercial lipoxygenase is used to generate the hydroperoxides, the above acids appear in about 2:1 proportions, respectively. The products of the action of the enzyme on the hydroperoxides were stabilized for analysis by reduction with H(2) and LiAIH(4). Trimethylsilyl derivatives of reduced products were analyzed by combined gas-liquid chromatography-mass spectrometry. Specific deuterium labeling permitted definite location of the oxo functions. (18)O(2) labeling experiments showed that the oxygens of both the oxo and the hydroxyl functions were derived from the hydroperoxide. Retention of both oxygens suggests that the reaction proceeds through a cyclic epiperoxide followed by a ketohydroxy-forming rearrangement. No products of hydroperoxide isomerase were found in reactions catalyzed by the crude enzyme from alfalfa seeds.  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