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.
Platelets, when stirred with 3 U thrombin/10(9) platelets, produced significant quantities of palmitoyllysophosphatidic acid (2.17 ng/10(9) platelets), stearoyllysophosphatidic acid (2.11 ng/10(9) platelets), and arachidonoyllysophosphatidic acid (1.06 ng/10(9) platelets). When platelets were pretreated with 100 microM of the phospholipase A2 inhibitor U10029A, there was a significant decrease in thrombin-stimulated production of stearoyllysophosphatidic acid (to 0.16 ng/10(9) platelets), while arachidonoyllysophosphatidic acid production was unchanged. U10029A concomitantly increased thrombin-stimulated production of stearoyl-containing phosphatidic acid species (primarily stearoylarachidonoylphosphatidic acid) from 5.99 to 9.71 ng/10(9) platelets. The results are consistent with the concept that stearoyllysophosphatidic acid production in platelets occurs via phospholipase A2 degradation of phosphatidic acid.  相似文献   

2.
A microbial isolate, Flavobacterium sp. strain DS5, produces 10-ketostearic acid (10-KSA) from oleic acid in 85% yield. This is the first report on this type of reaction catalyzed by a Flavobacterium strain. The product was purified to give white, plate-like crystals melting at 79.2°C. The optimum time, pH, and temperature for the production of 10-KSA are 36 h, 7.5, and 30°C, respectively. A small amount of 10-hydroxystearic acid (10-HSA) (about 10% of the amount of the main product, 10-KSA) is also produced during the bioconversion. 10-KSA is not further metabolized by strain DS5 and accumulates in the medium. In contrast to growing cells, a resting-cell suspension of strain DS5 produces 10-HSA and 10-KSA in a ratio of 1:3. The crude cell extract obtained from ultrasonic disruption of the cells converts oleic acid mainly to 10-HSA (10-HSA/10-KSA ratio, 97:3). This result strongly suggested that oleic acid is converted to 10-KSA via 10-HSA. Enzymes catalyzing the hydration and secondary alcohol dehydrogenation are cell associated. Product 10-HSA from strain DS5 is 66% enantiomeric excess in the 10(R) form. The oleic acid conversion enzyme(s) reacts with unsaturated fatty acids in the order oleic acid > palmitoleic acid > linoleic acid > linolenic acid > γ-linolenic acid > myristoleic acid.  相似文献   

3.
Jasmonic acid was identified from Mimosa pudica L. plants by mass spectrometry, high performance liquid chromatography and thin layer chromatography. Effects of authentic jasmonic acid on pulvinule movement and transpiration of the pinnae were compared with those of abscisic acid. Jasmonic acid and abscisic acid each at 10−5 M inhibited both auxin- and light-induced opening of the pulvinules. A closure-inducing activity of jasmonic acid at 10−4 M was greater than that of abscisic acid at 10−4 M. Pinnae transpiration was reduced by 10−5 M abscisic acid but not by 10−4 M jasmonic acid.  相似文献   

4.
Our previous study indicated that oleic acid prevented apoptotic cell death induced by trans10, cis12 (t10, c12)-conjugated linoleic acid in rat hepatoma dRLh-84 cells. The intracellular mechanism of action oleic acid is still unknown. Here, we showed that p38 mitogen-activated protein kinase (MAPK) inhibition using its specific inhibitor SB203580 cancelled the ameliorative effect of oleic acid on the cytotoxicity of t10, c12-conjugated linoleic acid. In addition, SubG1 cell population analysis showed that p38 MAPK played an essential role in the prevention of apoptotic cell death by oleic acid. In fact, p38 phosphorylation level was upregulated in cells treated with oleic acid irrespective of t10, c12-conjugated linoleic acid stimulation. Interestingly, t10, c12-conjugated linoleic acid increased intracellular triglyceride accumulation. However, oleic acid completely inhibited this effect. These observations indicated the involvement of blockade of a p38 MAPK pathway in the ameliorative effect of oleic acid on apoptosis induced by t10, c12-conjugated linoleic acid.  相似文献   

5.
Linoleic acid oxidation catalyzed by lipoxygenase (lipoxidase) activity in extracts of defatted corn germ does not terminate in the product, linoleic acid hydroperoxide, unless the lipoxygenase is first partially purified. If purification is not attempted, the hydroperoxide product exists only as a barely detectable intermediate in the synthesis of three products. One of these was identified as 9-hydroxy-10-oxo-cis-12-octadecenoic acid formed from the hydroperoxide by the enzyme, linoleate hydroperoxide isomerase. Another product, 13-hydroxy-10-oxo-trans-11-octadecenoic acid, is believed to be formed by an isomerase also. The third product was the linoleate ester of one of the hydroxy-oxo-fatty acids, 9-(cis-9,cis-12-octadecadienoyl)-10-oxo-cis-12-octadecenoic acid. It is not known if the synthesis of the ester is enzyme-catalyzed. When a mixture of 13-hydroperoxy-cis-9,trans-11-octa-decadienoic acid and 9-hydroperoxy-trans-10,cis-12-octa-decadienoic acid from soybean lipoxygenase oxidation of linoleic acid was used as a substrate, 13-hydroxy-12-oxo-cis-9-octadecenoic acid and 9-hydroxy-12-oxo-trans-10-octadecenoic acid were formed as the major products of catalysis by linoleate hydroperoxide isomerase(s) from corn. Smaller quantities of 9-hydroxy-10-oxo-cis-12-octadecenoic acid and 13-hydroxy-10-oxo-trans-11-octadecenoic acid were also formed.  相似文献   

6.
Four new aristololactams have been isolated from Aristolochia argentina. The evidence indicates them to be 10-amino-3-hydroxymethyl-2,4-dimethoxyphenanthrene-1-carboxylic acid lactam, 10-amino-3-hydroxymethyl-2,4,6-trimethoxyphenanthrene-1-carboxylic acid lactam, 10-amino-2-hydroxy-4-methoxyphenanthrene-1,3-dicarboxylic acid lactam and 10-amino-2-hydroxy-4,6-dimethoxyphenanthrene-1,3-dicarboxylic acid lactam.  相似文献   

7.
The ileal apical sodium bile acid cotransporter participates in the enterohepatic circulation of bile acids. In patients with primary bile acid malabsorption, mutations in the ileal bile acid transporter gene (Slc10a2) lead to congenital diarrhea, steatorrhea, and reduced plasma cholesterol levels. To elucidate the quantitative role of Slc10a2 in intestinal bile acid absorption, the Slc10a2 gene was disrupted by homologous recombination in mice. Animals heterozygous (Slc10a2+/-) and homozygous (Slc10a2-/-) for this mutation were physically indistinguishable from wild type mice. In the Slc10a2-/- mice, fecal bile acid excretion was elevated 10- to 20-fold and was not further increased by feeding a bile acid binding resin. Despite increased bile acid synthesis, the bile acid pool size was decreased by 80% and selectively enriched in cholic acid in the Slc10a2-/- mice. On a low fat diet, the Slc10a2-/- mice did not have steatorrhea. Fecal neutral sterol excretion was increased only 3-fold, and intestinal cholesterol absorption was reduced only 20%, indicating that the smaller cholic acid-enriched bile acid pool was sufficient to facilitate intestinal lipid absorption. Liver cholesteryl ester content was reduced by 50% in Slc10a2-/- mice, and unexpectedly plasma high density lipoprotein cholesterol levels were slightly elevated. These data indicate that Slc10a2 is essential for efficient intestinal absorption of bile acids and that alternative absorptive mechanisms are unable to compensate for loss of Slc10a2 function.  相似文献   

8.
Automatic recording of the frequency of feeding ‘bites’ was used to evaluate the effects of several organic acids (citric, metacectonic, lactic, acetic, and oxalic) on the stimulatory feeding behavior of Tilapia nilotica. Some of these acids are added to food stocks to retard spoilage. The results showed that citric acid at a concentration of 10?2 to 10?6 m , metacetonic acid at 10?4 to 10?6 m , and lactic acid at 10?2 to 10?5 m stimulated feeding. Fish tended to avoid metacetonic acid at 10?3 m and acetic acid at 10?3 m . Acetic acid at 10?5 m and oxalic acid at 10?6 m had no significant effects on fish feeding.  相似文献   

9.
通过规模化液态深层发酵获得灵芝发酵产物,采用多种硅胶色谱柱层析及重结晶的方式,从中分离得到10个化合物。通过核磁、质谱等波谱分析,鉴定出这些化合物均属于含羟基或酮基的不饱和脂肪酸类化合物,分别为(9S,10R,11E,13R)-9,10,13-trihydroxyoctadec-11-enoic acid(1)和(9S,10R,11E,13S)-9,10,13-trihydroxyoctadec-11-enoic acid(2)的混合物、12S*,13S*-dihydroxy-9-oxo-10(E)- octadecenoic acid(3)、9R*,10R*-dihydroxy-13-oxo-11(E)-octadecenoic acid(4)、12S*,13R*-dihydroxy- 9-oxo-10(E)-octadecenoic acid(5)、9S*,10R*-dihydroxy-13-oxo-11(E)-octadecenoic acid(6)、10(S)-hydroxy-8(Z)-octadecenoic acid(7)、12-oxooctadeca-8,10-dienoic acid(8)、9,12-dihydroxy-10-eicosenoic acid(9)和9-oxooctadeca-10,12-dienoic acid(10)。这些化合物均为首次从灵芝发酵产物中获得,且具有不同程度的体外抗肿瘤活性。其中,化合物8和化合物10对L1210细胞增殖抑制的IC50值分别为13.00μmol/L和16.88μmol/L,对K562细胞增殖亦有良好的抑制效果,是具有抗肿瘤潜力的天然产物。  相似文献   

10.
10-Hydroxystearic acid seems to be widely distributed in nature: Bacteria generate it by hydroxylation of oleic acid, but it was found also as constituent of plants, in cancer cell cultures and in mammalian tissue homogenates. Investigation of 10-hydroxystearic acid, obtained from mammalian tissue homogenates, revealed its identity with that of bacteria. Thus not 10-hydroxystearic acid is widely distributed in nature but its producers: bacteria. When biological material is processed in aqueous media, lipases are activated, these cleave membrane phospholipids. Thus liberated oleic acid is the substrate for widespread bacteria which are introduced into the media when the work up procedure is done in not sterile surrounding. The bacteria transform then oleic acid to 10R-hydroxystearic acid.  相似文献   

11.
(10L)- and (10D)-[1-14C, 10-3H]5,8,11,14,17-eicosapentaenoic acids were synthesized to investigate mechanistic and stereochemical aspects of leukotriene biosynthesis. Experiments with mastocytoma cells showed that a hydrogen is stereospecifically eliminated from C-10 during the conversion of eicosapentaenoic acid to leukotriene C5. The hydrogen lost has the pro-S (D) configuration. 5-Hydroxy-6,8,11,14,17-eicosapentaenoic acid, formed in the same experiments, was enriched in tritium when the (10D), but not when the (10L), isomer of labeled eicosapentaenoic acid was used. This indicates that oxygenation of the acid at C-5 occurred before the elimination of hydrogen and suggests that removal of the pro-S hydrogen at C-10 in 5-hydroperoxy-6,8,11,14,17-eicosapentaenoic acid initiates its transformation to trans-5(S),6(S)-oxido-7,9-trans-11,14,17-cis-eicosapentaenoic acid (leukotriene A5).  相似文献   

12.
R Labeque  L J Marnett 《Biochemistry》1988,27(18):7060-7070
Reaction of 10-hydroperoxyoctadec-8-enoic acid (10-OOH-18:1) (50 microM) with hematin (0.5 microM) in sodium phosphate buffer containing Tween 20 (200 microM) generates 10-oxooctadec-8-enoic acid, 10-oxodec-8-enoic acid (10-oxo-10:1), and 10-hydroxyoctadec-8-enoic acid in relative yields of 79, 4, and 17%, respectively. The product profile and relative distribution are unaffected by 1 mM butylated hydroxyanisole. Approximately 5% of the hydroperoxide isomerizes from the 10- to the 8-position. 10-Oxo-10:1 most likely arises via beta-scission of an intermediate alkoxyl radical to the aldehyde and the n-octyl radical. To test this, 10-hydroperoxyoctadeca-8,12-dienoic acid was reacted with hematin under identical conditions. 10-Oxooctadeca-8,12-dienoic acid, 10-oxodec-8-enoic acid, and 10-hydroxyoctadeca-8,12-dienoic acid are formed in relative yields of 50, 45, and 5%, respectively. The product ratios are constant with time and hydroperoxide to catalyst ratio and unaffected by inclusion of phenolic antioxidants. The higher yield of 10-oxo-10:1 from 10-OOH-18:2 compared to 10-OOH-18:1 is due to the higher rate of beta-scission of the intermediate alkoxyl radical from the former to the resonance-stabilized octenyl radical. Two products of reaction of the 2-octenyl radical with O2, octenal and octenol, were detected in 10% yield relative to 10-oxo-10:1. Inclusion of 7,8-dihydroxy-7,8-dihydrobenzo[a]pyrene (BP-7,8-diol) led to epoxidation by both 10-OOH-18:1 and 10-OOH-18:2. Studies with isotopically labeled hydroperoxide or O2 indicated approximately 65% of the epoxide oxygen was derived from O2 and 35% from hydroperoxide oxygen, consistent with the involvement of peroxyl free radicals as the oxidizing agents. The available evidence indicates that hematin reduces the fatty acid hydroperoxides homolytically to alkoxyl radicals that are oxidized to ketones, reduced to alcohols, or undergo beta-scission to aldehydes. Carbon radicals generated during these reactions couple to O2, generating peroxyl free radicals that epoxidize BP-7,8-diol. The smaller percentage of epoxidation that results from hydroperoxide oxygen may arise from oxidation of the hydroperoxide group to peroxyl radicals or from heterolytic cleavage of the hydroperoxide to alcohol and an iron-oxo complex.  相似文献   

13.
Plant phenols as in vitro inhibitors of glutathione S-transferase(s)   总被引:3,自引:0,他引:3  
Ellagic acid, a commonly occurring plant phenol, was shown to be a potent in vitro inhibitor of GSH-transferase(s) activity. Other plant phenols such as ferrulic acid, caffeic acid and chlorogenic acid also showed a concentration dependent inhibition of GSH-transferase(s) activity. The I50 values of ellagic acid, caffeic acid, chlorogenic acid and ferrulic acid were 8.3 X 10(-5)M, 14.0 X 10(-5)M, 20.0 X 10(-5)M and 22.0 X 10(-5)M respectively, suggesting that ellagic acid is the most potent inhibitor of all the four studied plant phenols. At 55 microM concentration of ellagic acid, a significant inhibition (35-47%) was observed on GSH-transferase activity towards CDNB, p-nitrobenzyl chloride and 1,2-epoxy-3-(p-nitrophenoxy)propane as substrates. Ellagic acid inhibited GSH-transferase(s) activity in a non-competitive manner with respect to CDNB while with respect to GSH it inhibited the enzyme activity in a competitive manner. Other phenolic compounds purpurogallin , quercetin, alizarin and monolactone also showed a concentration dependent inhibition of the enzyme activity with a I50 of 0.8 X 10(-5)M, 1.0 X 10(-5)M, 8.0 X 10(-5)M and 16.0 X 10(-5)M respectively. These inhibitors of GSH-transferase(s) activity should be useful in studying the in vitro enzyme mediated reactions of exogenous and endogenous compounds.  相似文献   

14.
Deuterium NMR spectra for a series of selectively deuterated substrates and inhibitors in the presence of lipoxygenase-1 (EC 1.13.11.12) are presented. Extrapolation of the 2H NMR line widths yielded transverse relaxation rates for the bound inhibitors [2H21]dodecanoic acid (protonated at the 2,2-position), [2,2-2H]dodecanoic acid, and [12,12,12-2H]dodecanoic acid which are 1/T2,bd = 5.0 X 10(3), 1.12 X 10(4), and 1.16 X 10(3) s-1, respectively. The substrates [9,10,12,13-2H]linoleic acid and [11,11-2H]linoleic acid had 1/T2,bd = 8.2 X 10(3) and 7.95 X 10(3) s-1, respectively. Kinetic measurements established Ki = 1.5 X 10(-3) M for dodecanoic acid (lauric acid) inhibition of lipoxygenase when the substrate is linoleic acid (Km = 2.6 X 10(-5) M). Lipoxygenase, with Mr 102,000, is predicted to have a rotational correlation time tau c - 1.2 X 10(-7) s, yielding a 1/T2,bd = 1.56 X 10(4) s-1 for tightly bound ligand. Hence, the correlation times of the selectively deuterated inhibitors indicate internal motions are present in the bound species.  相似文献   

15.
Both ascorbic acid and copper were strong prooxidants in the oxidation of linoleate in a buffered (pH 7.0) aqueous dispersion at 37 degrees C. Minimum concentrations at which catalytic activity was detected were 1.3 x 10(-7) m for copper and 1.8 x 10(-6) m for ascorbic acid. For concentrations up to 10(-3) m, the increase in rate of oxidation with increase in concentration of catalyst was greater for ascorbic acid than for copper. Ascorbic acid had maximum catalytic activity at 2.0 x 10(-3) m, but was still prooxidant at the highest concentration tested (5.0 x 10(-2) m). Dehydroascorbic acid was a weaker prooxidant than ascorbic acid. Further degradation products of ascorbic acid were not prooxidant. In early stages of the oxidation autocatalytic behavior was observed with copper, but not with ascorbic acid. Ascorbic acid functioned as a true catalyst, i.e., it accelerated the reaction but it was not oxidized simultaneously with the linoleate. It is proposed that the dehydroascorbic acid radical initiates the linoleate oxidation reaction.  相似文献   

16.
Adrenic acid (docosatetraenoic acid), an abundant fatty acid in the vasculature, is produced by a two-carbon chain elongation of arachidonic acid. Despite its abundance and similarity to arachidonic acid, little is known about its role in the regulation of vascular tone. Gas chromatography/mass spectrometric analysis of bovine coronary artery and endothelial cell lysates revealed arachidonic acid concentrations of 2.06 +/- 0.01 and 6.18 +/- 0.60 microg/mg protein and adrenic acid concentrations of 0.29 +/- 0.01 and 1.56 +/- 0.16 microg/mg protein, respectively. In bovine coronary arterial rings preconstricted with the thromboxane mimetic U-46619, adrenic acid (10(-9)-10(-5) M) induced concentration-related relaxations (maximal relaxation = 83 +/- 4%) that were similar to arachidonic acid relaxations. Adrenic acid relaxations were blocked by endothelium removal and the K(+) channel inhibitor, iberiotoxin (100 nM), and inhibited by the cyclooxygenase inhibitor, indomethacin (10 microM, maximal relaxation = 53 +/- 4%), and the cytochrome P-450 inhibitor, miconazole (10 microM, maximal relaxation = 52 +/- 5%). Reverse-phase HPLC and liquid chromatography/mass spectrometry isolated and identified numerous adrenic acid metabolites from coronary arteries including dihomo (DH)-epoxyeicosatrienoic acids (EETs) and DH-prostaglandins. DH-EET [16,17-, 13,14-, 10,11-, and 7,8- (10(-9)-10(-5) M)] induced similar concentration-related relaxations (maximal relaxations averaged 83 +/- 3%). Adrenic acid (10(-6) M) and DH-16,17-EET (10(-6) M) hyperpolarized coronary arterial smooth muscle. DH-16,17-EET (10(-8)-10(-6) M) activated iberiotoxin-sensitive, whole cell K(+) currents of isolated smooth muscle cells. Thus, in bovine coronary arteries, adrenic acid causes endothelium-dependent relaxations that are mediated by cyclooxygenase and cytochrome P-450 metabolites. The adrenic acid metabolite, DH-16,17-EET, activates smooth muscle K(+) channels to cause hyperpolarization and relaxation. Our results suggest a role of adrenic acid metabolites, specifically, DH-EETs as endothelium-derived hyperpolarizing factors in the coronary circulation.  相似文献   

17.
12-Oxo-trans-10-dodecenoic acid (trans-10-ODA) is an oxidation product of polyunsaturated fatty acids in plant tissues. The structural similarity of trans-10-ODA and traumatic acid, a compound considered to be a wound hormone, suggested that trans-10-ODA might be a precursor of traumatic acid. Both trans-10-ODA and traumatic acid were active in the Wehnelt bean assay. The results were more consistent with trans-10-ODA than with traumatic acid. Cucumber (Cucumis sativus L. var. National Pickling) hypocotyls also showed a growth increase following treatment with trans-10-ODA, which suggested that trans-10-ODA has a more general influence on plant development than previously ascribed to traumatic acid.  相似文献   

18.
Plant roots have important roles not only in absorption of water and nutrients, but also in stress tolerance such as desiccation, salt, and low temperature. We have investigated stress-response proteins from rice roots using 2-dimensional polyacrylamide-gel electrophoresis and found a rice protein, RO-292, which was induced specifically in roots when 2-week-old rice seedlings were subjected to salt and drought stress. The full-length RO-292 cDNA was cloned, and was determined to encode a protein of 160 amino acid residues (16.9 kDa, pI 4.74). The deduced amino acid sequence showed high similarity to known rice PR10 proteins, OsPR10a/PBZ1 and OsPR10b. RO-292 mRNA accumulated rapidly upon drought, NaCl, jasmonic acid and probenazole, but not by exposure to low temperature or by abscisic acid and salicylic acid. The RO-292 gene was also up-regulated by infection with rice blast fungus. Interestingly, induction was observed almost exclusively in roots, thus we named the gene RSOsPR10 (root specific rice PR10). The present results indicate that RSOsPR10 is a novel rice PR10 protein, which is rapidly induced in roots by salt, drought stresses and blast fungus infection possibly through activation of the jasmonic acid signaling pathway, but not the abscisic acid and salicylic acid signaling pathway.  相似文献   

19.
The conversion of linoleic acid into 10-hydroxy-12(Z)-octadecenoic acid by whole cells of Stenotrophomonas nitritireducens as an isolated bacterium was optimized, and the optimal temperature, pH, and cell and substrate concentrations were 30 degrees C, 7.5, and 20 and 20 g/L, respectively. Under these conditions, whole cells in a bioreactor produced 15 g/L 10-hydroxy-12(Z)-octadecenoic acid in 2 h of reaction time without detectable byproducts. Using 2 g/L linoleic acid, the cells produced 1.92 g/L 10-hydroxy-12(Z)-octadecenoic acid. These are the highest concentration and yield of 10-hydroxy-12(Z)-octadecenoic acid ever reported.  相似文献   

20.
To achieve demineralization of crab shell waste by chemical and biological treatments, lactic acid and lactic acid bacterium were applied. In 5.0 and 10% lactic acid, pH rapidly decreased from 6.8 to 4.2 and from 4.5 to 2.4 at day 3, respectively, and thereafter the pH remained at an almost constant level. In a 10% lactic acid bacterium inoculum, pH lowered to 4.6 at day 5. Relative residual ash content rapidly decreased to 49.1 and 16.4% in 5 and 10% lactic acid treatments, respectively, for the initial 12 h. In 2.5, 5 and 10% lactic acid bacterium inoculums, relative residual ash content rapidly decreased to 55.2, 40.9 and 44.7%, respectively, on the first day. Residual dry masses were 76.4, 67.8 and 46.6% in 2.5, 5 and 10% lactic acid treatments, respectively, for the initial 12 h. After a one-time exchange of the lactic acid solution, in the 5.0% lactic acid treatment, residual dry mass rapidly decreased from 66.0 to 41.4%. In 2.5, 5 and 10% lactic acid bacterium inoculums, residual dry masses decreased to 67.6, 57.4 and 59.6% respectively, on the first day. Protein contents after demineralization ranged from 51.3–54.7% in the chemical treatments and decreased to 32.3% in the lactic acid fermentation process. A negative relationship was shown between pH and demineralization rate in lactic acid and lactic acid bacterium treatments. These results suggest that lactic acid fermentation can be an alternative for demineralization of crab shells, even though the rate and efficiency of the demineralization is lower than the chemical treatment.  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