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.
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   

2.
The diterpene acid content in 10 species of Helianthus has been investigated. Ent-12,16-cyclokauranoic acid, isolated from H. annuus, is converted into a series of 12,16-cyclogibberellins by cultures of Gibberella fujikuroi, mutant B1-41a, and 12,16-cyclogibberellins A9, and A12 have been isolated. Ent-12β-acetoxykaurenoic acid and ent-13(S)-angeloxyatisenoic acid have been isolated from H. decapetalus; the metabolism of ent-13(S)-hydroxyatisenoic acid and atisenoic acid by B1-41a is also described.  相似文献   

3.
An inducible enzyme catalysing the hydrolysis of (+)-usnic acid to (+)-2-desacetylusnic acid and acetic acid has been purified 150-fold from the mycelium of Mortierella isabellina grown in the presence of (+)-usnic acid. Purification was achieved by treatment with protamine sulfate, (NH4)2SO4 fractionation, negative adsorption on alumina Cγ gel and hydroxylapatite followed by chromatography on DEAE-cellulose and Sephadex G-200. The elution pattern from a Sephadex G-200 column indicated a MW of ca 7.6 × 104 for the enzyme. The apparent Km value for (+)-usnic acid at the pH optimum (pH 7) was 4.0 × 10?5 M. The enzyme was specific for (+)-usnic acid and inactive towards (?)-usnic acid, (+)-isousnic acid or certain phloracetophenone derivatives. Its activity was enhanced in the presence of divalent metal ions such as Co2+, Ni2+, Mn2+, Mg2+ and Zn2+.  相似文献   

4.
Both ellagic and gallic acids non competitively inhibited guaiacol oxidation by turnip peroxidase. The Ki values were 3 and 26 μm for ellagic and gallic acid respectively. Enzymatic oxidation of gallic acid by the isolated major turnip peroxidase was characterized with respect to spectral behaviour, affinity constant and pH effect. The Km for H2O2 and gallic acid are 2.5 and 8.0 mM for turnip peroxidase. The pH optimum for gallic acid oxidation is about 6.5 and the rate constant k4 decreased with the increase of pH in presence of both guaiacol and Gallic acid. When the gallic acid oxidation products were subjected to chromatographic analysis, it was found to be converted mainly to ellagic and an unknown quinone.  相似文献   

5.
Quinolinic acid phosphoribosyltransferase (EC 2.4.2.19) was purified 3600-fold from rat liver and 280-fold from rat brain. Kinetic analyses (Km = 12 μM for the substrate quinolinic acid and Km 23 μM for the cosubstrate phosphoribosylpyrophosphate), physicochemical properties of the purified enzymes, inhibition by phthalic acid (Ki = 1.4 μM) and molecular weight determination (Mr 160 000 for the holoenzyme, consisting of five identical 32 kDa subunits) indicated the structural identity of quinolinic acid phosphoribosyltransferase from the two rat tissues. This was further confirmed immunologically, using antibodies raised against purified rat liver quinolinic acid phosphoribosyltransferase. Rat quinolinic acid phosphoribosyltransferase differs in several aspects from quinolinic acid phosphoribosyltransferase isolated from other organisms. The purified enzyme will prove a useful tool in the examination of a possible role of quinolinic acid in cellular function and/or dysfunction.  相似文献   

6.
Decomposition of aqueous solutions of gibberellic acid on autoclaving   总被引:1,自引:0,他引:1  
R.J. Pryce 《Phytochemistry》1973,12(3):507-514
A qualitative and quantitative analysis of the decomposition of unbuffered and buffered (pH 3–8) aqueous solutions of gibberellic acid (GA3) on autoclaving is recorded. The identified products, which vary in composition with pH, are iso-GA3 (II), iso-GA3 hydroxy acid (III), gibberellenic acid (IV), allogibberic acid (V), epiallogibbe detected after autoclaving in all cases. The identified products, in all cases, account for not less than 95% of the decomposition product, Dehydroallogibberic acid has not previously been recorded as an aqueous decomposition product of GA3 and its biological activity in the lettuce hypocotyl test is recorded.  相似文献   

7.
α-Aminoisobutyric acid is actively transported into yeast cells by the general amino acid transport system. The system exhibits a Km for α-aminoisobutyric acid of 270 μM, a Vmax of 24 nmol/min per mg cells (dry weight), and a pH optimum of 4.1–4.3. α-Aminoisobutyric acid is also transported by a minor system(s) with a Vmax of 1.7 nmol/min per mg cells. Transport occurs against a concentration gradient with the concentration ratio reaching over 1000:1 (in/out). The α-aminoisobutyric acid is not significantly metabolized or incorporated into protein after an 18 h incubation. α-Aminoisobutyric acid inhibits cell growth when a poor nitrogen source such as proline is provided but not with good nitrogen sources such as NH4+. During nitrogen starvation α-aminoisobutric acid strongly inhibits the synthesis of the nitrogen catabolite repression sensitive enzyme, asparaginase II. Studies with a mutant yeast strain (GDH-CR) suggest that α-aminoisobutyric acid inhibition of asparaginase II synthesis occurs because α-aminoisobutyric acid is an effective inhibitor of protein synthesis in nitrogen starved cells.  相似文献   

8.
The dehydrogenation of substituted 3,5-cyclohexadiene-1,2-diol-1-carboxylic acids by dihydrodihydroxybenzoic acid dehydrogenases from benzoate grown cells of Alcaligenes eutrophus and Pseudomonas sp. B 13 and 3 -chlorobenzoate grown cells of the latter organism was examined. No significant differences (Km and Vrel values) were detected for the enzymes from both organisms. The same dihydrodihydroxybenzoic acid dehydrogenase is formed in Pseudomonas sp. B13 during growth on benzoate as well as on 3-chlorobenzoate. The lower turnover rates of 3- and 5-chlorodihydrodihydroxybenzoic acid compared to dihydrodihydroxybenzoic acid are counterbalanced by an increase in specific activity. With the exception of 4-substituted dihydrodihydroxybenzoic acids exhibiting relative high Km values, only slight sterical and electronic substituent effects are evident. Reaction rates were never reduced to a critical level.  相似文献   

9.
The aim of the present study was to identify the enzymes in human liver catalyzing hydroxylations of bile acids. Fourteen recombinant expressed cytochrome P450 (CYP) enzymes, human liver microsomes from different donors, and selective cytochrome P450 inhibitors were used to study the hydroxylation of taurochenodeoxycholic acid and lithocholic acid. Recombinant expressed CYP3A4 was the only enzyme that was active towards these bile acids and the enzyme catalyzed an efficient 6α-hydroxylation of both taurochenodeoxycholic acid and lithocholic acid. The Vmax for 6α-hydroxylation of taurochenodeoxycholic acid by CYP3A4 was 18.2 nmol/nmol P450/min and the apparent Km was 90 μM. Cytochrome b5 was required for maximal activity. Human liver microsomes from 10 different donors, in which different P450 marker activities had been determined, were separately incubated with taurochenodeoxycholic acid and lithocholic acid. A strong correlation was found between 6α-hydroxylation of taurochenodeoxycholic acid, CYP3A levels (r2=0.97) and testosterone 6β-hydroxylation (r2=0.9). There was also a strong correlation between 6α-hydroxylation of lithocholic acid, CYP3A levels and testosterone 6β-hydroxylation (r2=0.7). Troleandomycin, a selective inhibitor of CYP3A enzymes, inhibited 6α-hydroxylation of taurochenodeoxycholic acid almost completely at a 10 μM concentration. Other inhibitors, such as α-naphthoflavone, sulfaphenazole and tranylcypromine had very little or no effect on the activity. The apparent Km for 6α-hydroxylation of taurochenodeoxycholic by human liver microsomes was high (716 μM). This might give an explanation for the limited formation of 6α-hydroxylated bile acids in healthy humans. From the present results, it can be concluded that CYP3A4 is active in the 6α-hydroxylation of both taurochenodeoxycholic acid and lithocholic acid in human liver.  相似文献   

10.
Ascorbalamic acid (C9H13NO8) was isolated from Brassica olerocea L. MS study of various methylated derivatives suggested a structure (Ia) derivable by CC coupling of C-3 of alanine with C-2 of ascorbic acid, followed by lactone → lactam rearrangement. Other derivatives provided supporting evidence, as did study of the reaction of L-3-chloroalanine with L-ascorbic acid in vitro. On treatment with hot 6 M HCl, ascorbalamic acid yielded L-aspartic acid and 3-(2-furoyl)alanine. For identification of the latter, DL-3-(2-furoyl)alanine and its N-2,4-dinitrophenyl and N-acetyl methyl ester derivatives were synthesized. Unlike ascorbigens, ascorbalamic acid is probably present in the living plant. It seemed to be present in all crucifers examined, but to have a capricious distribution in other orders. During permethylation, rearrangements of ester groups were observed, both with ascorbalamic acid and with pyrrolidonecarboxylic acid as a model.  相似文献   

11.
MUFAs are unsaturated FAs with one double bond and are derived from endogenous synthesis and dietary intake. Accumulating evidence has suggested that plasma and erythrocyte MUFA levels are associated with cardiometabolic disorders, including CVD, T2D, and metabolic syndrome (MS). Previous genome-wide association studies (GWASs) have identified seven loci for plasma and erythrocyte palmitoleic and oleic acid levels in populations of European origin. To identify additional MUFA-associated loci and the potential functional variant at each locus, we performed ethnic-specific GWAS meta-analyses and trans-ethnic meta-analyses in more than 15,000 participants of Chinese and European ancestry. We identified novel genome-wide significant associations for vaccenic acid at FADS1/2 and PKD2L1 [log10(Bayes factor) ≥ 8.07] and for gondoic acid at FADS1/2 and GCKR [log10(Bayes factor) ≥ 6.22], and also observed improved fine-mapping resolutions at FADS1/2 and GCKR loci. The greatest improvement was observed at GCKR, where the number of variants in the 99% credible set was reduced from 16 (covering 94.8 kb) to 5 (covering 19.6 kb, including a missense variant rs1260326) after trans-ethnic meta-analysis. We also confirmed the previously reported associations of PKD2L1, FADS1/2, GCKR, and HIF1AN with palmitoleic acid and of FADS1/2 and LPCAT3 with oleic acid in the Chinese-specific GWAS and the trans-ethnic meta-analyses. Pathway-based analyses suggested that the identified loci were in unsaturated FA metabolism and signaling pathways. Our findings provide novel insight into the genetic basis relevant to MUFA metabolism and biology.  相似文献   

12.
The hypothesis that ambient CO2 levels determine the end-products of energy metabolism excreted by Hymenolepis diminuta was tested by incubating the parasite in a range of CO2 concentrations and measuring internal concentrations of adenine nucleotides and the excretion of organic acids. The strain of H. diminuta used was found to excrete mainly lactic acid and acetic acid. Succinic acid production was generally less than 5–10% of the total. At high CO2 concentrations, the rate of excretion of lactic acid decreased while that of succinic acid increased, which conforms with the hypothesis. Acetic acid excretion did not vary significantly over the range of CO2 concentrations used. Other results did not support the hypothesis. High CO2 levels reduced the total amounts of acids excreted and the rate of succinic acid excretion was so small as to be ineffective in preventing the accumulation of H+ ions. When present in the incubation medium, succinic acid was taken up by H. diminuta. Lactic and acetic acid excretion was always sufficient to limit the accumulation of H+ ions. The conditions of incubation were shown not to be responsible for the low rates of succinic acid excreted. Incubation conditions and metabolic end-products were found to affect the rates of excretion of organic acids. There is thus a need, in work of this nature, to regulate and specify experimental conditions and to stipulate the strain of parasite used. The hypothesis was rejected and it was suggested that the energy metabolism of parasitic helminths is adapted to fluctuating O2 and CO2 tensions.  相似文献   

13.
Extracts of maize leaves catalyzed the interconversion of meso-diaminopimelic acid its L-isomer. Three observations support the existence of this epimerase activity: (i) detection of the reversible interconversion of L-diaminopimelic acid and meso-diaminopimelic acid by paper chromatography after incubation of either isomer with extract; (ii) formation of [14C]CO2 from L-[14C]diaminopimelic acid in an incubation mix containing meso-diaminopimelic acid decarboxylase; and (iii) inhibition of [14C]CO2 evolution from L-diaminopimelic acid by unlabeled meso-diaminopimelic acid. The demonstration of the diaminopimelic acid epimerase lends support to the occurrence in plants of the complete diaminopimelic acid pathway for biosynthesis of lysine as it occurs in Escherichia coli and most bacteria.  相似文献   

14.
Some properties of a preparation of an enzyme, lunularic acid decarboxylase, from the liverwort Conocephalum conicum are described. The enzyme is normally bound and could be solubilized with Triton X-100; at least some of the bound decarboxylase activity appears to be associated with chloroplasts. For lunularic acid the enzyme has Km 8.7 × 10?5 M (pH 7.8 and 30°). Some substrate analogues have been tested but no other substrate was found. Pinosylvic acid is a competitive inhibitor for the enzyme, Ki 1.2 × 10?4 M (pH 7.8 and 30°). No product inhibition was observed. Lunularic acid decarboxylase activity has also been observed with a cell-free system from Lunularia cruciata.  相似文献   

15.
In vitro studies on the fatty acid metabolism of the epimastigotes and trypomastigotes of Trypanosoma cruzi showed the following: (1) Trypomastigotes demonstrated the ability to convert labeled palmitic acid to CO2. Epimastigotes either did not convert this fatty acid to CO2 or did so at a very low rate. (2) Trypomastigotes incorporated palmitic acid into neutral lipids, but, at a rate less than that of the epimastigotes. (3) While epimastigotes readily incorporated palmitic acid into phospholipid lipids, trypomastigote forms seemed to lack this ability.  相似文献   

16.
The PUFAs include many bioactive lipids. The microbial metabolism of C18 PUFAs is known to produce their bioactive isomers, such as conjugated FAs and hydroxy FAs, but there is little information on that of C20 PUFAs. In this study, we aimed to obtain anaerobic bacteria with the ability to produce novel PUFAs from C20 PUFAs. Through the screening of ∼100 strains of anaerobic bacteria, Clostridium bifermentans JCM 1386 was selected as a strain with the ability to saturate PUFAs during anaerobic cultivation. This strain converted arachidonic acid (cis-5,cis-8,cis-11,cis-14-eicosatetraenoic acid) and EPA (cis-5,cis-8,cis-11,cis-14,cis-17-EPA) into cis-5,cis-8,trans-13-eicosatrienoic acid and cis-5,cis-8,trans-13,cis-17-eicosatetraenoic acid, giving yields of 57% and 67% against the added PUFAs, respectively. This is the first report of the isolation of a bacterium transforming C20 PUFAs into corresponding non-methylene-interrupted FAs. We further investigated the substrate specificity of the biohydrogenation by this strain and revealed that it can convert two cis double bonds at the ω6 and ω9 positions in various C18 and C20 PUFAs into a trans double bond at the ω7 position. This study should serve to open up the development of novel potentially bioactive PUFAs.  相似文献   

17.
Extraction of Rudbeckia fulgida furnished 13αH-labd-8(17)-en-15-al-19-oic acid, two new C18-acids tentatively formulated as 9 (S*),12 (S*),13 (S*)-trihydroxyoctadeca-10(E),15 (Z)-dienoic acid and 9 (S*),12 (S*),13 (S*)-trihydroxyoctadec-10 (E)-enoic acid, several known C14-polyacetylenes and several flavone glycosides.  相似文献   

18.
The metabolism of several ring C and D-functionalized ent-kaur-16-en-19-oic acids by cultures of Gibberella fujikuroi, mutant B1-41a, to the corresponding derivatives of the normal fungal gibberellins (GAs) and ent-kaurenoids is described. A range of 12α- and 12β-hydroxyGAs and ent-kaurenoids are characterized by their mass spectra and GC Kovats retention indices. The mass spectral and GC data are used to identify the 12α-hydroxy derivatives of GA12, GA14, GA37 and GA4 (GA58), and of the 12β-hydroxy derivatives of ent-7α-hydroxy- and ent-6α, 7α-dihydroxykaurenoic acids, in seeds of Cucurbita maxima. Similarly the metabolites of GA9, formed in seeds of Pisum sativum and cultures of G.fujikuroi, mutant B1-41a, are identified as 12α-hydroxyGA9. ent-11β-Hydroxy- and ent-11-oxo-kaurenoic acids are metabolized by the fungus to the corresponding 11-oxygenated derivatives of the normal fungal ent-kaurenoids and some C20-GAs; no 11-oxygenated C19-GAs are formed. Grandiflorenic acid, 11β-hydroxygrandiflorenic acid, attractyligen and ent-15β-hydroxykaurenoic acid are metabolized to unidentified products.  相似文献   

19.
The metabolism of vulpinic acid by an unclassified soil micro-organism was studied. A new compound, 2,5-diphenyl-3-hydroxy-4-oxo-2-hexendioic acid (DHOHA) was isolated from the reaction mixture of a cell-free preparation and pulvinic acid. The existence of a hydrolase which catalyses the conversion of vulpinic acid to pulvinic acid was detected in cell-free preparation, and an inducible lactone hydrolase capable of converting pulvinic acid to DHOHA was purified 130-fold and characterized. This enzyme had a MW of ca 34 000, a Km for pulvinic acid at pH optimum (pH 7.0) less than 10 ? 6 M, pI = 5.0, and was inhibited by p-chloromercuriphenylsulfonate and diethylpyrocarbonate. The enzyme was highly specific for pulvinic acid. The initial degradative steps proposed for this organism are vulpinic acid → pulvinic acid → DHOHA.  相似文献   

20.
Abietic and dehydroabietic acid are interesting diterpenes with a highly diverse repertoire of associated bioactivities. They have, among others, shown antibacterial and antifungal activity, potentially valuable in the struggle against the increasing antimicrobial resistance and imminent antibiotic shortage. In this paper, we describe the synthesis of a set of 9 abietic and dehydroabietic acid derivatives containing amino acid side chains and their in vitro antimicrobial profiling against a panel of human pathogenic microbial strains. Furthermore, their in vitro cytotoxicity against mammalian cells was evaluated. The experimental results showed that the most promising compound was 10 [methyl N-(abiet-8,11,13-trien-18-yl)-d-serinate], with an MIC90 of 60 μg/mL against Staphylococcus aureus ATCC 25923, and 8 μg/mL against methicillin-resistant S. aureus, Staphylococcus epidermidis and Streptococcus mitis. The IC50 value for compound 10 against Balb/c 3T3 cells was 45 μg/mL.  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