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 966 毫秒
1.
Dehydration of 2-(2-arylethyl)-2-hydroxy-4-oxopentanoic acids 1 with hydrochloric acid/acetic acid, affords 3-(2-arylethyl)-5-hydroxy-5-methyl-2(5H)-furanones 4 . Compounds of type 1 and 4 represent suitable precursors for the formation of pyridazin-3-ones 7 as they smoothly react with hydrazine. A new series of s-triazolo[4,3-b]pyridazin-3-ones 12 and tetrazolo[1,5-b]pyridazines 15 are obtained from the 3-chloropyridazines 11 upon treatment with semicarbazide and sodium azide, respectively. Reaction of 11 with phenyl- acetyl-hydrazine provides 3-benzyl-6-phenyl-8-(2-phenyl-ethyl)-s-triazolo[4,3-b]pyridazine 13 via dehydrative cyclization of the intermediate 14 which was clarified to exhibit tautomeric equilibria between enol–hydrazine form A and keto–hydrazine form B by means of 1H and 13C NMR spectroscopy. Attempts to synthesize 3-alloxy-pyridazines 18 by reacting 11 with sodium alloxide afford N-allyl compounds 17 .  相似文献   

2.
Mary P. Carpenter 《Lipids》1974,9(6):397-406
The purpose of the study was to determine whether prostaglandins were present in mammalian testis, a tissue that has a large concentration of polyenoic fatty acids that are potential precursors of prostaglandins. Acid-soluble lipids of rat testis were extracted, purified, and fractionated by thin layer and column chromatographies.3H-Prostaglandins were added as internal reference standards to monitor recoveries and facilitate identification. Initial identification of prostaglandin species was done by chromatography. Further identification was done by elution of the prostaglandin zones followed by rechromatographies (both thin layer and column), measurements of UV absorption spectra, and by gas liquid chromatography. The results of these analyses indicate that prostaglandin E1, 11α,15(S)-dihydroxy-9-oxo-β-trans-prostenoic acid; prostaglandin E2, 11α-15(S)-dihydroxy-9-oxo-5-cis-13-trans-prostadienoic acid; and prostaglandin F1, 9α,11α,15(S)-trihydroxy-13-trans-prostenoic acid occur in rat testicular tissue and that prostaglandin F, 9α, 11α, 15(S)-trihydroxy-5-cis-13-trans-prostadienoic acid and prostaglandin E2, 11α,15(S)-dihydroxy-9-oxo-5-cis-13-trans-prostadienoic acid may be the primary species of this tissue. Prostaglandin B1, 15(S)-hydroxy-9-oxo-8(12),13-trans-prostadienoic acid and prostaglandin B2, 15(S)-hydroxy-9-oxo-5-cis,8(12),13-trans-prostatrienoic acid also were detected, and some evidence was obtained for the presence of prostaglandin metabolites.  相似文献   

3.
Luo Y  Yi J  Li B  Zhang G 《Lipids》2004,39(9):907-913
Novel ceramides, rel-(3S,4S,5S)-3-[(2R)-2-hydroxycosanoyl-hexacosanoylamino]-4-hydroxy-5-[(4Z)-tetradecane-4-ene]-2,3,4,5-tetrahydrofuran (1a-g), and a new glucoceramide, 1-O-β-d-glucopyranosyl-(2S,3S,4R,8E)-2-[(2R)-2-hydroxytetracosanoylamino]-1,3,4-octodecanetriol-8-ene (2) were isolated from the aqueous ethanolic extract of the roots of Incarvillea arguta, together with eight known compounds: β-sitosterol (3), oleanolic acid (4), ursolic acid (5), piperin (6), maslinic acid (7), β-sitosterol 6′-O-acyl-β-d-glucopyranoside (8), 8-epideoxyloganic acid (9), and plantarenaloside (10). Their structures were elucidated on the basis of spectral data including IR, MS, NMR [1H NMR, 13C NMR (distortionless enhancement by polarization transfer), 1H−1H COSY, heteronuclear multiplequantum coherence, and heteronuclear multiple-bond coherence correlations]. The relative configurations were established by nuclear Overhauser effect spectroscopy experiments and by comparison of the NMR spectral data and coupling constants with those already reported in the literature.  相似文献   

4.
4,8-Dimethyldecanal (4,8-DMD) is the aggregation pheromone produced by male red flour beetles (RFB), Tribolium castaneum. To elucidate the biosynthetic origin of 4,8-DMD, the following studies were performed: (1)effects of juvenile hormone (JH) III, and pathway inhibitors mevastatin, an inhibitor of the mevalonate pathway, and 2-octynoic acid, an inhibitor of the fatty acid pathway, were tested to determine whether 4,8-DMD is derived from the fatty acid pathway or the mevalonate pathway; (2) incorporation of 13C-labeled acetate, propionate, and mevalonolactone into 4,8-DMD was measured to directly determine the biosynthetic origin of 4,8-DMD; and (3) incorporation of deuterium-labeled precursors, including 2-methylbutanoate (C5D), 4-methylhexanoate (C7D), 2,6-dimethyloctanoate (C10D), and 4,8-dimethyldecanoate (C12D) was tested to determine whether 4,8-DMD is biosynthesized in the sequence AcPrAcPrAc (Ac; acetate, Pr; propionate). JH III was topically applied to males at various doses. Inhibitors and isotopically labeled substrates were administered orally by feeding the beetles flour treated with the substrates of interest, after which volatiles were collected from both sexes of RFBs. The amount of 4,8-DMD produced was significantly reduced with increasing doses of JH III. Also, 2-octynoic acid inhibited the production of 4,8-DMD, but mevastatin did not. Exposure of RFBs to [1-13C]acetate and [1-13C]propionate, but not [2-13C]mevalonolactone, resulted in incorporation of the labeled compounds into 4,8-DMD. RFBs fed flour treated with deuterium-labeled C5D, C10D, and C12D, but not C7D, incorporated these compounds into 4,8-DMD. The findings that the production of 4,8-DMD was inhibited by 2-octynoic acid but unaffected by mevastatin, combined with the high incorporation of [1-13C]acetate and [1-13C]propionate into 4,8-DMD and the incorporation of deuterated precursors, unambiguously demonstrated that 4,8-DMD is of fatty acid rather than terpene biosynthetic origin, and that the biosynthesis of 4,8-DMD proceeds in the sequence Ac-Pr-Ac-Pr-Ac.  相似文献   

5.
The additional mass associated with 13C in metabolic tracers may interfere with their metabolism. The comparative isomerization and biohydrogenation of oleic, [1-13C]oleic, and [U-13C]oleic acids by mixed ruminal microbes was used to evaluate this effect. The percent of stearic, cis-14 and- 15, and trans-9 to-16 18∶1 originating from oleic acid was decreased for [U-13C]oleic acid compared with [1-13C]oleic acid. Conversely, microbial utilization of [U-13C]oleic acid resulted in more of the 13C label in cis-9 18∶1 compared with [1-13C]oleic acid (53.7 vs. 40.1%). The isomerization and biohydrogenation of oleic acid by ruminal microbes is affected by the mass of the labeled tracer.  相似文献   

6.
Nε-(6-purinoyl)-L-lysine has been prepared by acylation of L-lysine copper complex or Nα-carbobenzoxy-L-lysine with 6-trichloromethylpurine followed by the removal of the protecting group. β-(5-uracilyl)-DL-alanine has been prepared by condensation of 5-chloro-methyluracil with diethyl acetamidomalonate, followed by acid hydrolysis.  相似文献   

7.
Laverroux S  Glasser F  Gillet M  Joly C  Doreau M 《Lipids》2011,46(9):843-850
In ruminants, cis and trans C18:1 isomers are intermediates of fatty acid transformations in the rumen and their relative amounts shape the nutritional quality of ruminant products. However, their exact synthetic pathways are unclear and their proportions change with the forage:concentrate ratio in ruminant diets. This study traced the metabolism of vaccenic acid, the main trans C18:1 isomer found in the rumen, through the incubation of labeled vaccenic acid with mixed ruminal microbes adapted to different diets. [1-13C]trans-11 C18:1 was added to in vitro cultures with ruminal fluids of sheep fed either a forage or a concentrate diet. 13C enrichment in fatty acids was analyzed by gas-chromatography-mass spectrometry after 0, 5 and 24 h of incubation. 13C enrichment was found in stearic acid and in all cis and trans C18:1 isomers. Amounts of 13C found in fatty acids showed that 95% of vaccenic acid was saturated to stearic acid after 5 h of incubation with the concentrate diet, against 78% with the forage diet. We conclude that most vaccenic acid is saturated to stearic acid, but some is isomerized to all cis and trans C18:1 isomers, with probably more isomerization in sheep fed a forage diet.  相似文献   

8.
To highlight the role of plants in traditional healing, the leaves and the stems of cultivated Triumfetta cordifolia were phytochemically studied yielding a new nor-ursane type (1), a new ceramide (2) and a new piperidinic ceramide derivative (3) named, respectively, 2α,19α-dihydroxy-3-oxo-23-nor-urs-12-en-28-oic acid, (2R)-2-hydroxy-N-[(2S,3S,4R,26E)-1,3,4-trihydroxy-26-triaconten-2-yl] tetradecanamide and (2R,8Z)-2-hydroxy-{(2S,3R,5R,6S)-3,5-dihydroxy-6-[(1E,5Z)-hexadeca-1,5-dienyl]-2-(β-d-glucopyranosyloxy)methyl piperidine-1-yl} tetracos-8-enamide (3). These were obtained together with lupeol (4), stigmasterol (5), 3-O-β-d-glucopyranoside of β-sitosterol (6), tormentic acid (7) from stems and heptadecanoic acid (8), β-carotene (9), oleanolic acid (10), and 24-hydroxytormentic acid (11) from leaves. The structures were determined on the basis of NMR data (1H-, 13C-, 2D-NMR analyses), mass spectrometry and confirmed by chemical transformations as well as comparison of spectral data with those reported in the literature. The FRAP method was used to evaluate the antioxidant activity of fractions collected from flash chromatography and isolated compounds. Among the fractions, four reduced FeIII-TPTZ to FeII-TPTZ while isolated pure compounds showed no activity.  相似文献   

9.
To develop a synthesis of tritiated cyclopropene fatty acids (CPFA), compounds that should prove useful for affinity labeling of desaturases in insect pheromone biosynthetic studies, a series of novel, selectively deuterated CPFA analogues was prepared and characterized. In methyl [16-2H]12,13-methylene-12-hexadecenoate, the incorporation of deuterium was achieved by treatment of the corresponding ω-chloro derivative with sodium borodeuteride in dimethylsulfoxide at 70°C for 24 h (67% yield) following conventional procedures. Alkylation of the tetrahydropyranyl derivative of 13-tridecynol in the presence of lithium diisopropylamide in tetrahydrofuran at −20°C with 1-chloro-3-iodopropane in hexamethylphosphoramide, followed by Jones oxidation of the crude product, yielded 16-chloro-12-hexadecynoic acid (54%), which was esterified to the corresponding methyl ester by treatment with potassium carbonate and methyl iodide in dimethylformamide. Treatment of this acetylenic ester with ethyldiazoacetate in the presence of activated copper-bronze as catalyst followed by hydrolysis in KOH solution at room temperature yielded 16-chloro-12,13-(carboxymethylene)-12-hexadecenoic acid. This diacid was treated with excess oxalyl chloride to give the corresponding diacyl chloride, which was decarbonylated in a diethyl ether solution with zinc chloride, and the cyclopropenium ions thus formed were added at −40°C to a methanolic sodium hydroxide solution of sodium borohydride to give methyl 16-chloro-12,13-methylene-12-hexadecenoate. Analogous procedures were followed to prepare methyl [17-2H]10,11-methylene-10-hexadecenoate, methyl [17-2H]11,12-methylene-11-hexadecenoate and methyl [17-2H]12,13-methylene-12-hexadecenoate from the corresponding diacids using sodium borodeuteride in the reduction of the cyclopropenium ions. Alternatively, methyl [2,2,3,3-2H4]hexadecynoate, prepared by reaction of methyl 2,11-hexadecadiynoate with magnesium in deuterated methanol at room temperature, was submitted to the above cyclopropenylation and reductive decarbonylation sequence to give methyl [2,2,3,3,17-2H5]-11,12-methylene-11-hexadecenoate. In summary, complementary methods for the selective incorporation of one to five deuterium atoms into cyclopropene fatty acids, at different sites, in moderate to high yields have been developed. The methods should easily be applicable to the preparation of the corresponding tritiated analogues.  相似文献   

10.
Ernst H. Oliw  Mats Hamberg 《Lipids》2019,54(9):543-556
Fusarium oxysporum f. sp. tulipae (FOT) secretes (+)-7-iso-jasmonoyl-(S)-isoleucine ((+)-JA-Ile) to the growth medium together with about 10 times less 9,10-dihydro-(+)-7-iso-JA-Ile. Plants and fungi form (+)-JA-Ile from 18:3n-3 via 12-oxophytodienoic acid (12-OPDA), which is formed sequentially by 13S-lipoxygenase, allene oxide synthase (AOS), and allene oxide cyclase (AOC). Plant AOC does not accept linoleic acid (18:2n-6)-derived allene oxides and dihydrojasmonates are not commonly found in plants. This raises the question whether 18:2n-6 serves as the precursor of 9,10-dihydro-JA-Ile in Fusarium, or whether the latter arises by a putative reductase activity operating on the n-3 double bond of (+)-JA-Ile or one of its precursors. Incubation of pentadeuterated (d5) 18:3n-3 with mycelia led to the formation of d5-(+)-JA-Ile whereas d5-9,10-dihydro-JA-Ile was not detectable. In contrast, d5-9,10-dihydro-(+)-JA-Ile was produced following incubation of [17,17,18,18,18-2H5]linoleic acid (d5-18:2n-6). Furthermore, 9(S),13(S)-12-oxophytoenoic acid, the 15,16-dihydro analog of 12-OPDA, was formed upon incubation of unlabeled or d5-18:2n-6. Appearance of the α-ketol, 12-oxo-13-hydroxy-9-octadecenoic acid following incubation of unlabeled or [13C18]-labeled 13(S)-hydroperoxy-9(Z),11(E)-octadecadienoic acid confirmed the involvement of AOS and the biosynthesis of the allene oxide 12,13(S)-epoxy-9,11-octadecadienoic acid. The lack of conversion of this allene oxide by AOC in higher plants necessitates the conclusion that the fungal AOC is distinct from the corresponding plant enzyme.  相似文献   

11.
The hydrocarbons (3Z,6Z,9Z)-3,6,9-octadecatriene (3Z,6Z,9Z-18:H) and (3Z,6Z,9Z)-3,6,9-nonadecatriene (3Z,6Z,9Z-19:H) constitute the pheromone of the winter moth, Erannis bajaria. These compounds belong to a large group of lepidopteran pheromones which consist of unsaturated hydrocarbons and their corresponding oxygenated derivatives. The biosynthesis of such hydrocarbons with an odd number of carbons in the chain is well understood. In contrast, knowledge about the biosynthesis of even numbered derivatives is lacking. We investigated the biosynthesis of 3Z,6Z,9Z-18:H by applying deuterium-labeled precursors to females of E. bajaria followed by gas chromatography–mass spectrometry analysis of extracts of the pheromone gland. A mixture of deuterium-labeled [17,17,18,18-2H4]-3Z,6Z,9Z-18:H and the unlabeled 3Z,6Z,9Z-18:H was obtained after topical application and injection of (10Z,13Z,16Z)-[2,2,3,3-2H4]-10,13,16-nonadecatrienoic acid ([2,2,3,3-2H4]-10Z,13Z,16Z-19:acid) or (11Z,14Z,17Z)-[3,3,4,4-2H4]-11,14,17-icosatrienoic acid ([3,3,4,4-2H4]-11Z,14Z,17Z-20:acid). These results are consistent with a biosynthetic pathway that starts with α-linolenic acid (9Z,12Z,15Z-18:acid). Chain elongation leads to 11Z,14Z,17Z-20:acid, which is shortened by α-oxidation as the key step to yield 10Z,13Z,16Z-19:acid. This acid can be finally reduced to an aldehyde and decarbonylated or decarboxylated to furnish the pheromone component 3Z,6Z,9Z-18:H. A similar transformation of 11Z,14Z,17Z-20:acid yields the second pheromone component, 3Z,6Z,9Z-19:H.  相似文献   

12.
Mosley EE  McGuire MA 《Lipids》2007,42(10):939-945
There is limited methodology available to quantitatively assess the activity of the Δ9-desaturase enzyme in vivo without chemically inhibiting the enzyme or using radioactively labeled substrates. The objective of these experiments was to develop methodology to determine the incorporation and desaturation of 13C-labeled fatty acids into milk lipids. In a preliminary experiment, 3.7 g [1-13C]myristic acid ([1-13C]14:0), 19.5 g [1-13C]palmitic acid ([1-13C]16:0), 20.0 g [1-13C]stearic acid ([1-13C]18:0) were combined and infused into the duodenum of a cow over 24 h. In a following experiment, 5.0 g [1-13C]14:0, 40.0 g [1-13C]16:0, and 50.0 g [1-13C]18:0 were infused into the abomasums of separate cows as a bolus over 20 min or continuously over 24 h. Milk fat was extracted using chloroform:methanol. Fatty acids were methylated, and fatty acid methyl esters (FAME) were converted to dimethyl disulfide derivatives (DMDS). The FAME and DMDS were analyzed by gas chromatography mass spectrometry. In the preliminary experiment, 13C enrichment in 14:0 but not 16:0 or 18:0 was observed. When dosage amounts were increased in the following experiment, peak enrichments from the bolus infusion were observed at 8 h. Enrichments for continuous infusion peaked at 16 h for 14:0 and 18:0, and at 24 h for 16:0. The Δ9-desaturase products of these fatty acids were estimated to be 90% of cis-9 14:1, 50% of cis-9 16:1, and 59% of cis-9 18:1. This study demonstrates that 13C-labeled fatty acids may be utilized in vivo to measure the activity of the Δ9-desaturase enzyme.  相似文献   

13.
Hamberg M 《Lipids》2011,46(2):201-206
Exposure of unsaturated fatty acids to singlet oxygen results in the formation of hydroperoxides. In this process, each double bond in the acyl chain produces two regioisomeric hydroperoxides having an (E)-configured double bond. Although such compounds are racemic, the hydrogen removal associated with the oxygenation may, a priori, take place antarafacially, suprafacially or stereorandomly. The present study describes the preparation of [11(S)-2H]linoleic acid by two independent methods and the use of this stereospecifically labeled fatty acid to reveal the hidden stereospecificity in singlet oxygenations of polyunsaturated fatty acids. It was found that linoleic acid 9(R)- and 13(S)-hydroperoxides formed from [11(S)-2H]linoleic acid both retained the deuterium label whereas the 9(S)- and 13(R)-hydroperoxides were essentially devoid of deuterium. It is concluded that polyunsaturated fatty acid hydroperoxides produced in the presence of singlet oxygen in e.g., plant leaves are formed by a reaction involving addition of oxygen and removal of hydrogen taking place with suprafacial stereochemistry. This result confirms and extends previous mechanistic studies of singlet oxygen-dependent oxygenations.  相似文献   

14.
Products of the Thermal Ene-Reaction of Unsaturated Fatty Compounds and Maleic Anhydride The thermal ene reaction of the methylates of unsaturated fatty acids 10-undecenoic acid, oleic acid and (E)-10-eicosendioic acid with maleic anhydride was carried out at 190°C. The mono addition products were isolated and their stereochemistry was deduced from 1HNMR- and 13C-NMR-spectroscopy.  相似文献   

15.
Our objective was to develop a suitable probe to study metabolism of polyunsaturated fatty acids by 13C nuclear magnetic resonance (NMR) in the suckling rat pup. [3-13C] γ-Linolenic acid was chemically synthesized, and a 20 mg (Experiment 1) or 5 mg (Experiment 2) dose was injected into the stomachs of 6–10-day-old suckling rat pups that were then killed over a 192 h (8 d) time course. 13C NMR showed that 13C in γ-linolenate peaked in liver total lipids by 12-h post-dosing and that [5-13C]-arachidonic acid peaked in both brain and liver total lipids 48–96 h post-dosing. 13C enrichment in brain γ-linolenic acid was not detected by NMR, but gas chromatography-combustion-isotope ratio mass spectrometry showed that its mass enrichment in brain phospholipids at 48–96 h post-dosing was 1–2% of that in brain arachidonic acid. 13C was present in liver and brain cholesterol and in perchloric acid-extractable water-soluble metabolites in the brain, liver and carcass. We conclude that low but measurable amounts of exogenous γ-linolenic acid do access the suckling rat brain in vivo. The slow time course of [5-13C] arachidonic acid appearance in the brain suggests most of it was probably transported there after synthesis elsewhere, probably in the liver. Some carbon from γ-linolenic acid is also incorporated into lipid products other than n−6 long-chain polyunsaturated fatty acids.  相似文献   

16.
During our ongoing project on the biosynthesis of R-(+)-octane-1,3-diol the metabolism of linoleic acid was investigated in stored apples after injection of [1-14C]-, [9,10,12,13-3H]-, 13C18- and unlabeled substrates. After different incubation periods the products were analyzed by gas chromatography-mass spectroscopy (MS), high-performance liquid chromatography-MS/MS, and HPLC-radiodetection. Water-soluble compounds and CO2 were the major products whereas 13(R)-hydroxy- and 13-keto-9(Z),11(E)-octadecadienoic acid, 9(S)-hydroxy-and 9-keto-10(E),12(Z)-octadecadienoic acid, and the stereoisomers of the 9,10,13- and 9,12,13-trihydroxyoctadecenoic acids were identified as the major metabolites found in the diethyl ether extracts. Hydroperoxides were not detected. The ratio of 9/13-hydroxy- and 9/13-keto-octadecadienoic acid was 1∶4 and 1∶10, respectively. Chiral phase HPLC of the methyl ester derivatives showed enantiomeric excesses of 75% (R) and 65% (S) for 13-hydroxy-9(Z),11(E)-octadecadienoic acid and 9-hydroxy-10(E),12(Z)-octadecadienoic acid, respectively. Enzymatically active homogenates from apples were able to convert unlabeled linoleic acid into the metabolites. Radiotracer experiments showed that the transformation products of linoleic acid were converted into (R)-octane-1,3-diol. 13(R)-Hydroxy-9(Z), 11(E)-octadecadienoic acid is probably formed in stored apples from 13-hydroperoxy-9(Z),11(E)-octadecadienoic acid. It is possible that the S-enantiomer of the hydroperoxide is primarily degraded by enzymatic side reactions, resulting in an enrichment of the R-enantiomer and thus leading to the formation of 13(R)-hydroxy-9(Z),11(E)-octadecadienoic acid.  相似文献   

17.
Four hydrophobic amidine containing monomers N-(pyridin-2-yl)methacrylamide (3), 5-(pyridin-2-ylamino)pentyl methacrylate (7), S-5-(isopropenylcarbonyloxy)pentyl-N,N-diphenylthiourea (9) and methyl-5-guanidino-2-methacrylamidopentanoate (12) were synthesized and copolymerized with a water soluble co-monomer N,N-dimethylacrylamide (13) or N-isopropylacrylamide (22). The second component was a copolymer containing methacrylic acid (19) and 13/22. By combination of these two components the formation of amidinium-carboxylate interactions was determined by dynamic light scattering (DLS), 1H-NMR spectroscopy and rheology measurements.  相似文献   

18.
Mechanism of Long Chain Monoenoic Fatty Acids Acting on the Energy Metabolism of Heart The oxidation of 1-14C-erucic (C22:1) and 1-14C-nervonic (C24:1) acids was studied in comparison to 1-14C-palmitic and -oleic acids in isolated rat and pig heart mitochondria. After mitochondrial incubation with the albumin-bound fatty acids only small amounts of 14CO2 developed from the oxidation of the long chain monoenoic acids as compared to palmitic or oleic acid. The slow down of the oxidation rate was more pronounced in rat than in pig heart mitochondria. The oxidation of palmitic or oleic acid was not found to be inhibited by the C20–C24-monoenoic acids, whereas palmitic or oleic acid inhibited the oxidation of erucic acid competitively. From present findings an idea may be developed of the interference on fatty acid metabolism in heart muscle by erucic and other long chain monoenoic acids.  相似文献   

19.
Catalyzed by 10−5 M ionic iron in 80% ethanol,N-acetylcysteine added to linoleic acid hydroperoxide, forming a thiobond. Reaction of a specific isomer of the hydroperoxide, 13-hydroperoxy-trans-11,cis-9-octadecadienoic acid, andN-acetylcysteine, forms a number of products, of which two were identified as addition compounds. One addition product was 9-S-(N-acetylcysteine)-13-hydroxy-10-ethoxy-trans-11-octadecenoic acid, and the other was 9-S-(N-acetylcysteine)-10,13-dihydroxy-trans-11-octadecenoic acid. Presented at the AOCS 49th Annual Fall Meeting, Cincinnati, September–October, 1975.  相似文献   

20.
N-tert-Butyl-N-[1-diethylphosphono-(2,2-dimethylpropyl)]nitroxide (DEPN) as β-phosphonylated nitroxide was prepared using tert-butylamine, pivalaldehyde and phosphonic acid diethyl ether as starting materials. This nitroxide revealed a reversible redox peak in cyclic voltammetry at +0.81 V versus Ag/AgCl, and showed electrocatalysis for the oxidation of alcohol. A preparative electrocatalytic oxidation of 4-methylbenzyl alcohol on the nitroxide yielded 98% 4-methylbenzaldehyde by 10 h of electrolysis. The current efficiency, turnover number and pseudo-first-order kinetic constant of the reaction were 96.5%, 39 and 0.95×10−5 s−1, respectively.  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