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 46 毫秒
1.
After the formation of dianions of a carboxylic acid with lithium diisopropylamide, oxygen was bubbled into the solution to produce 2-hydroperoxy acid. Then the reaction mixture was acidified with a 2 N HCl solution and subsequently elevated to 50 °C to afford the aldehyde with the loss of one carbon atom. Even saturated (C10–C20) and unsaturated (C18:1) carboxylic acids were converted into the odd aldehydes (C9–C19, C17:1) in high yields. This conversion was found to be an efficient method for the preparation of carboxylic acids (Cn) to one-carbon degraded aldehydes (Cn-1) via 2-hydroperoxy acids.  相似文献   

2.
In order to know the substrate specificity in a hydrocarbon utilizing bacterium, the following materials were examined: n-alkanes, n-alkenes, monohydric alcohols, aldehydes, monobasic carboxylic acids, dihydric alcohols and dibasic carboxylic acids.

It was found that dibasic carboxylic acids were well utilized, and a great deal of l-glutamic acid was accumulated from them. Then suberic acid, which is C8 dibasic carboxylic acid, was compared with n-dodecane in the effects of thiamine, penicillin, C/N ratio and substrate concentration on l-glutamic acid accumulation and cell growth.  相似文献   

3.
The pyrolysis reactions of soybean oils have been studied. The pyrolytic products were analyzed by GC–MS and FTIR and show the formation of olefins, paraffins, carboxylic acids and aldehydes. Several kinds of catalysts were compared. It was found that the amounts of carboxylic acids and aldehydes were significantly decreased by using base catalysts such as Na2CO3 and K2CO3. The low acid value pyrolytic products showed good cold flow properties and good solubility in diesel oil at low temperature. The results presented in this work have shown that the pyrolysis of soybean oils generates fuels that have chemical composition similar to petroleum based fuels.  相似文献   

4.
Homogenates of tomato fruits catalysed the enzymic conversion of linoleic and linolenic acids (but not oleic acid) to C6 aldehydes in low (3–5%) molar yield. Hexanal was formed from linoleic acid; cis-3-hexenal and smaller amounts of trans-2-hexenal were formed from linolenic acid. With the fatty acids as substrates, the major products were fatty acid hydroperoxides (50–80% yield) and the ratio of 9- to 13-hydroperoxides as isolated from an incubation with linoleic acid was at least 95:5 in favour of the 9-hydroperoxide isomer. When the 9- and 13-hydroperoxides of linoleic acid were used as substrates with tomato homogenates, the 13-hydroperoxide was readily cleaved to hexanal in high molar yield (60%) but the 9-hydroperoxide isomer was not converted to cleavage products. Properties of the hydroperoxide cleavage system are described. The results indicate that the C6 aldehydes are formed from C18 polyunsaturated fatty acids in a sequential enzyme system involving lipoxygenase (which preferentially oxygenates at the 9-position) followed by a hydroperoxide cleavage system which is, however, specific for the 13-hydroperoxy isomers.  相似文献   

5.
When treated with acetic or propionic acid vapours, intact pre-climacteric Golden Delicious apples started ripening prematurely. Experiments with [2?14C]propionic acid showed that a small part of the added acid is transformed into [14C]ethylene, and that ripening probably begins in these circumstances as a result of the artificial increase in the endogenous ethylene level. It may be that in unripe apples the small amount of evolved ethylene is mostly derived from simple organic acids, so that the moment at which its triggering concentration is reached depends on the available supply of acids. As this, in turn, is a function of the state of maturity of the fruit, there must be a direct relationship between the start of ripening and the degree of maturity. Thus, even in very unripe apples, catabolic processes are at work or potentially present because of the modified β–oxidation of [1?14C]propionic acid observed.  相似文献   

6.
A series or γ- and δ-lactones could be found in the thermal oxidative products of normal saturated acids, aldehydes, and alcohols (C9, C10, and C12, respectively) heated at 180°C in the presence of 0.1% KMnO4. Their lactones were identified by gas chromatography, infrared spectroscopy, and mass spectroscopy. And they could be detected also in the volatile compounds occurred by heating of C10 acid, aldehyde, and alcohol mixed with pork fat. So it was expected that lactones in meat fat flavor described in the earlier papers could be secondary products converted from saturated acids, aldehydes, and alcohols formed by oxidative degradation of meat fats. This process was presumed to be one of the mechanisms of the lactone formation.

It was discussed that lactones might be derived through mono or dihydroperoxides of acids, aldehydes, and alcohols.  相似文献   

7.
The molecular composition of lipids in three samples of leafy mosses (Aulacomnium palustre, Warnstorfia fluitans, and Calliergon giganteum) has been determined. The revealed acyclic compounds included normal and isoprenoid alkanes, isoprenoid alkenes, normal and isoprenoid ketones, carboxylic acids and their esters, alcohols, and aldehydes. Among cyclic compounds, bi-, tri- and tetracyclic polycycloaromatic hydrocarbons (PAHs), bicyclic and pentacyclic terpenoids, steroids and tocopherols have been observed. The identified organic compounds consisted mainly of carbocyclic acids and n-alkanes with the prevalence of C27 homologues. A. palustre is characterized by a reduced content of isoprenoid compounds, alcohols, and ketones, while the content of unsaturated acids, pentacyclic terpenoids, and aldehydes is rather heightened. A. palustre differs from W. fluitans and C. giganteum in the steroid composition and contains eremophylene, a sesquiterpenoid, which is absent in the mosses of the family Amblystegiaceae. Compared to C. giganteum, W. fluitans has a higher content of lycopadiene, carboxylic acids, n-alkanes, phyt-2-ene, aldehydes, esters, squalene, diploptene, α-tocopherol, and triphenyl phosphates.  相似文献   

8.
Abstract— Brains of human fetuses III, V, VI, VII months) newborn and infants (3, 7 and 13 months old) were investigated and the contents of total lipid, neutral- and phospholipid fractions were estimated. Fatty acids as well as fatty aldehydes of the phosphatides were analysed by gas chromatography. The results showed, that during this period of development the C16-compounds in the fatty acid and aldehyde fractions decrease, while the total C18-derivatives increase. However, the C18-monoenoic fatty acids decrease from the third fetal month until birth and increase during myelination. The same pattern was found for the C18-monoenoic aldehydes. The amounts of C20- and C22-polyenoic fatty acids were relatively constant. Only trace amounts of aldehydes with chain lengths other than C16- and C18-saturated and C18-monoenoic comnniinds were found.  相似文献   

9.
It is shown that the production of ethylene by whole apples,discs of peel prepared from them, and extracts prepared fromthe discs and supplied with linolenic acid, increases simultaneouslyduring the development of the respiration climacteric in apples.As the climacteric peak is reached, the ability of the extractsto produce ethylene declines and this is associated with a rapidloss of ethylene-producing activity on short term (up to 24h) ageing of the peel discs from which the extracts were obtained. It is suggested that the gmn.ll ethylene production by extracts(with linolenic acid) from pre-climacteric fruits, which arenot themselves evolving sufficient ethylene for its detectionin the ambient atmosphere, may be explained on the grounds thateven before the climacterio the fruit is producing small amountsof ethylene within the tissue; the tissue, as the climactericphase is approached, becomes more sensitive to ethylene andautocatalysed production of the gas then stimulates the fruitinto the respiration climacteric.  相似文献   

10.
An engineered reversal of the β-oxidation cycle was exploited to demonstrate its utility for the synthesis of medium chain (6–10-carbons) ω-hydroxyacids and dicarboxylic acids from glycerol as the only carbon source. A redesigned β-oxidation reversal facilitated the production of medium chain carboxylic acids, which were converted to ω-hydroxyacids and dicarboxylic acids by the action of an engineered ω-oxidation pathway. The selection of a key thiolase (bktB) and thioesterase (ydiI) in combination with previously established core β-oxidation reversal enzymes, as well as the development of chromosomal expression systems for the independent control of pathway enzymes, enabled the generation of C6–C10 carboxylic acids and provided a platform for vector based independent expression of ω-functionalization enzymes. Using this approach, the expression of the Pseudomonas putida alkane monooxygenase system, encoded by alkBGT, in combination with all β-oxidation reversal enzymes resulted in the production of 6-hydroxyhexanoic acid, 8-hydroxyoctanoic acid, and 10-hydroxydecanoic acid. Following identification and characterization of potential alcohol and aldehyde dehydrogenases, chnD and chnE from Acinetobacter sp. strain SE19 were expressed in conjunction with alkBGT to demonstrate the synthesis of the C6–C10 dicarboxylic acids, adipic acid, suberic acid, and sebacic acid. The potential of a β-oxidation cycle with ω-oxidation termination pathways was further demonstrated through the production of greater than 0.8 g/L C6–C10 ω-hydroxyacids or about 0.5 g/L dicarboxylic acids of the same chain lengths from glycerol (an unrelated carbon source) using minimal media.  相似文献   

11.
We have cloned, sequenced, and expressed the gene for a unique ATP- and NADPH-dependent carboxylic acid reductase (CAR) from a Nocardia species that reduces carboxylic acids to their corresponding aldehydes. Recombinant CAR containing an N-terminal histidine affinity tag had Km values for benzoate, ATP, and NADPH that were similar to those for natural CAR, and recombinant CAR reduced benzoic, vanillic, and ferulic acids to their corresponding aldehydes. car is the first example of a new gene family encoding oxidoreductases with remote acyl adenylation and reductase sites.  相似文献   

12.
Fatty aldehydes are an important group of fragrance and flavor compounds that are found in different fruits and flowers. A biotechnological synthesis of fatty aldehydes based on Escherichia coli cells expressing an α-dioxygenase (αDOX) from Oryza sativa (rice) is presented. α-Dioxygenases are the initial enzymes of α-oxidation in plants and oxidize long and medium-chain C n fatty acids to 2-hydroperoxy fatty acids. The latter are converted to C n − 1 fatty aldehydes by spontaneous decarboxylation. Successful expression of αDOX in E. coli was proven by an in vitro luciferase assay. Using resting cells of this recombinant E. coli strain, conversion of different fatty acids to the respective fatty aldehydes shortened by one carbon atom was demonstrated. The usage of Triton X 100 improves the conversion rate up to 1 g aldehyde per liter per hour. Easy reuse of the cells was demonstrated by performing a second biotransformation without any loss of biocatalytic activity.  相似文献   

13.
Cytochrome P450 (P450) 2A6 activates nitrosamines, including N,N-dimethylnitrosamine (DMN) and N,N-diethylnitrosamine (DEN), to alkyl diazohydroxides (which are DNA-alkylating agents) and also aldehydes (HCHO from DMN and CH3CHO from DEN). The N-dealkylation of DMN had a high intrinsic kinetic deuterium isotope effect (Dkapp ∼ 10), which was highly expressed in a variety of competitive and non-competitive experiments. The Dkapp for DEN was ∼3 and not expressed in non-competitive experiments. DMN and DEN were also oxidized to HCO2H and CH3CO2H, respectively. In neither case was a lag observed, which was unexpected considering the kcat and Km parameters measured for oxidation of DMN and DEN to the aldehydes and for oxidation of the aldehydes to the carboxylic acids. Spectral analysis did not indicate strong affinity of the aldehydes for P450 2A6, but pulse-chase experiments showed only limited exchange with added (unlabeled) aldehydes in the oxidations of DMN and DEN to carboxylic acids. Substoichiometric kinetic bursts were observed in the pre-steady-state oxidations of DMN and DEN to aldehydes. A minimal kinetic model was developed that was consistent with all of the observed phenomena and involves a conformational change of P450 2A6 following substrate binding, equilibrium of the P450-substrate complex with a non-productive form, and oxidation of the aldehydes to carboxylic acids in a process that avoids relaxation of the conformation following the first oxidation (i.e. of DMN or DEN to an aldehyde).  相似文献   

14.
The cutins of fruits and leaves of four apple cultivars have been analysed using TLC, GLC and GC-MS. They are similarly composed of saturated, monounsaturated and diunsaturated fatty, hydroxy-fatty and epoxyhydroxy-fatty acids. The most abundant monomers are 18-hydroxyoctadeca-9,12-dienoic, 10,16-dihydroxyhexadecanoic, 9,10-epoxy-18-hydroxyoctadec-12-enoic, 9,10-epoxy-18-hydroxyoctadecanoic and 9,10,18-trihydroxyoctadecanoic acids. The fruit cutins have high contents of epoxides (35–40%) and unsaturated components ( > 40%) and C18 compounds predominate over C16. The leaf cutins contain smaller amounts of unsaturated components than the fruits and higher proportions of C16 compounds. The adaxial leaf cutin differs in composition from the abaxial. 10,16-Dihydroxyhexadecanoic and 9,10-epoxy-18-hydroxoctadecanoic acids are the major constituents (each ca. 30%) of the adaxial leaf cutin and 10,16-dihydroxyhexadecanoic acid (55–65%) predominates in the abaxial.  相似文献   

15.
The course of biosynthesis of fatty acids in the seeds of winter rape (Brassica napus L. ssp.oleifera, f.biennis cv. T?ebí?ská) was investigated. After the termination of flowering seed samples were taken at five intervals, the seeds were divided into 4 fractions according to size, and their weight, water content, oil content and fatty acid composition were determined. The oil content was found to increase in all size categories with time, with the exception of a minute drop when complete maturity is reached. Larger seeds contained more oil. The fatty acid composition changes with time in the individual size fractions almost continuously. The same holds for differences between seed sizes of the same sample. The main change in oil composition consists in the decrease of C18 acids in favour of C22 acids. Greatest decrements during maturation were found with oleic acid, less with linoleic acid. In absolute amounts the quantity of all synthesized acids rises, the greatest rise being observed with C22 acids (i.e. predominantly erucic acid). It follows from the mean rates of synthesis of the individual groups (C16, C18, C20, C22) of fatty acids that the fraction of C22 rate of synthesis increases, while that of the C18 acids decreases with the same speed. The results indicate that the fatty acid synthesis is most intense during the second half of seed maturation, the main role being played by accelerating the synthesis of higher acids, especially of erucic acid.  相似文献   

16.
Nocardia 107-332, a soil isolate, oxidizes short-chain alkyl-substituted cyclic hydrocarbons to cyclic acids while growing on n-alkanes. Cyclic acids are produced also from relatively long-chain alkyl-substituted cyclics such as n-nonylbenzene or n-dodecylbenzene which alone support growth in a mineral-salts medium. ω-Oxidation of the alkyl substituents is followed by β-oxidation. It is of particular interest that cyclic acids such as cyclohexaneacetic and phenylacetic with C2 residual carboxylic acid substituents are resistant to further oxidation by the nocardia but cyclic acids with C1 or C3 substituents are readily oxidized and utilized for growth.

The specificity of microbial oxidations is demonstrated by the conversion of p-isopropyltoluene (p-cymene) to p-isopropylbenzoic acid in n-alkane, growth-supported nocardia cultures.

  相似文献   

17.
Epicuticular wax from mature plants of Sorghum bicolor SD-102 was compared with that from panicles and seedlings of the same variety at the fourth-fifth leaf stage of growth. The composition of wax from SD-102 panicles was quite different from that of mature leaf blades and sheaths. Free fatty alcohols were the dominant class of wax from SD-102 seedlings and C32 was the major homologue of alcohols and aldehydes. For comparison purposes, the epicuticular waxes from whole plants of two other S. bicolor varieties, Alliance A and Martin A, grown up to the tasseling stage, have been analysed. The major wax components were free fatty acids. The typical chain lengths of aldehydes, free alcohols and free fatty acids were C28 and C30.p-Hydroxybenzaldehyde was not a wax component of the studied varieties of sorghum.  相似文献   

18.
M.J.K. Macey 《Phytochemistry》1974,13(8):1353-1358
Two different mutations in Brassica oleracea, gl5 and gl4 have been re-investigated using acetate-1-14C labelling in an attempt to define more closely the nature of the genetic blocks to wax synthesis. It has been found that gl5 is a mutation which blocks elongation in the Step C28–C30. The mutation gl4 exhibits no elongation block and could be blocked in the decarboxylation Step C30–C29. 0·1 mM TCA supplied in the culture solution of cauliflower seedlings affected the leaf surface by producing a glossy appearance similar to that induced by gl3 and gl4. At this concentration growth was not inhibited and the appearance of the plants was normal except for the surface wax. The amount of surface wax produced was about 40% of that in untreated seedlings on a leaf area basis. Slight, but significant changes in wax composition were noted, mainly involving a reduction in C30 acids and aldehydes, a slight reduction (33–29%) in alkane content, and a marked difference in chain length composition of the alkanes with C27 increased relative to C29. Over a range of concentrations from 0·1–1 mM, TCA inhibited incorporation of label from acetate-1-14C into C30 acids and aldehydes more than into C28 at concentrations 0·4–0·8 mM while label tended to accumulate in C24 and C26 acids; thus elongation C28–C30 was especially sensitive to TCA. TCA also inhibited incorporation into primary alcohols and esters almost as much as into C29 compounds. In spite of relatively specific effects on incorporation of label into longer chain lengths, the resulting block to C30 synthesis is not sufficient to make much difference to the overall rate of C29 synthesis. Both results of analysis of wax from whole plants and experiments with tissue slices in vitro indicated that the effect of TCA in reducing the glaucousness of the leaf surface is a combination of overall reduction of wax synthesis together with slight but significant changes in wax composition.  相似文献   

19.
为探讨植物在机械损伤后C6-C10醛类化合物的释放机理, 及C6-C10醛类化合物对叶片光系统II (PSII)的影响, 以樟树(Cinnamomum camphora)为材料, 采用动态顶空气体循环法和热脱附/气相色谱/质谱联用技术(TDS-GC-MS), 对樟树幼苗叶片损伤后释放的C6-C10醛类化合物进行采集与分析, 并同步测定了脂氧合酶活性和损伤叶片的叶绿素荧光动力学参数。结果表明: 樟树幼苗叶片损伤后, 其挥发性有机化合物中己醛、庚醛、辛醛、壬醛和癸醛的释放量比损伤前分别增加了2.47、0.99、1.34、0.91和28.38倍(p < 0.01); 同时新增4种醛类化合物, 分别是: 2-己烯醛、2,4-己二烯醛、(E)-2-辛烯醛和(E)-2-壬烯醛。脂氧合酶活性比损伤前增加1.2倍(p < 0.01)。PSII单位反应中心复合体吸收的能量和被核心捕获的能量分别比损伤前下降12.8%和9.8% (p < 0.01)。单位面积反应中心的数量、电子传递量子产额、捕获激子能导致电子传递效率和叶片性能指数分别比损伤前增加23.3%、24.4%、22.6%和82.7% (p < 0.01)。损伤24 h后, 醛类化合物的种类、释放量、脂氧合酶活性及叶片叶绿素荧光动力学参数基本恢复到损伤前水平。说明机械损伤使PSII供体侧受损、脂氧合酶活性升高, 致使C6-C10醛类化合物大量释放, 樟树幼苗通过增加单位面积反应中心的数量来提高光合效率应对胁迫。  相似文献   

20.
There are three mycolic acid homologues with C22-, C24- and C26-α-units in Mycobacterium. In order to reveal the composition and distribution of these homologues in each subclass and molecular species of mycolic acids and to compare them with the composition of constitutive non-polar fatty acids (free and bound forms), we have separated non-polar fatty acids and each subclass of mycolic acids from 21 mycobacterial species by thin-layer chromatography, and analyzed non-polar fatty acid methyl esters by gas chromatography (GC) and the cleavage products of methyl mycolate by pyrolysis GC. We further performed mass chromatographic analysis of trimethylsilyl (TMS) ether derivatives of mycolic acid methyl esters by monitoring [B-29]+ ions (loss of CHO from the α-branched-chain structure of mycolic acids) of m/z 426, 454 and 482 which are attributed to C22-, C24- and C26-α-units of TMS ether derivatives of methyl mycolates, respectively, (Kaneda, K. et al, J. Clin. Microbiol. 24: 1060-1070, 1986). By pyrolysis GC, C22:0, C24:0 and C26:0 fatty acid methyl esters generated by the C2-C3 cleavage of C22-, C24- and C26-α-unit-containing mycolic acid methyl esters, respectively, were detected. Their proportion was almost the same among subclasses of mycolic acids in every Mycobacterium and also similar to the proportion of constitutive non-polar C22:0, C24:0 and C26:0 fatty acids. By mass chromatography, the composition and distribution of C22- and C24-α-unit-containing homologues were revealed to be similar between α- and α'-mycolic acids in every Mycobacterium. We further analyzed in detail M. vaccae and demonstrated that the mass chromatogram of C22-α-unit-containing homologue was analogous in shape to that of the C24-α-unit-containing one, with the latter mass chromatogram being up-shifted from the former by two carbon numbers, in every subclass of α-, α'-, keto and dicarboxy mycolic acids. The present study suggests that the compositions of three homologues of both mycolic acids and constitutive non-polar fatty acids, which are characteristic to each mycobacterial species, may reflect the proportion of the amount of free C22:0, C24:0 and C26:0 fatty acids synthesized in the cell. It is further demonstrated that intermolecular condensation of two fatty acids which become α- and β-units of mycolic acids will occur independently of the carbon chain length or kinds of polar moieties of fatty acids.  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