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.
Metathesis of Unsaturated Fatty Acid Esters – a Simple Approach to Long-Chained Dicarboxylic Acids Long-chained, symmetric unsaturated C18-, C20- und C26-dicarboxylic acid esters are easily accessible from natural fatty acid esters by metathesis in a two-step procedure. In the first step unsaturated fatty acid esters are cleaved by metathesis with ethylene. Pure oleic acid ester or fatty acid methylesters, produced from high-oleic sunflower oil or from rapeseed oil by transesterification with methanol, are converted to 1-decene and 9-decenoic acid methylester. From erucic acid methylester 1-decene and 13-tetradecenoic acid methylester are achieved. Using our newly developed high efficient catalytic system B2O3-Re2O7/Al2O3-SiO2 + SnBu4 conversion rates of 86 – 96% are obtained and the terminally unsaturated fatty acid esters are isolated in yields of 61 – 83%. In the second step 9-decenoic and 13-tetradecenoic acid methylester as well as 10-undecenoic methylester, which is commercially obtained by pyrolysis of ricinoleic acid ester, are converted to 9-octadecenedioic-, 13-hexacosenedioic- and 10-eicosenedioic acid dimethylester, by which process ethylene is eliminated. The conversion rates are 89 – 99% and the dicarboxylic acid dimethylesters are isolated in yields of 77 – 84%.  相似文献   

2.
A novel solid acid was prepared from petroleum coke by KOH chemical activation and concentrated H2SO4 sulfonation. The solid acid was characterized by XRD, FT-IR and solid-state NMR. The characterization results show that the chemical activation and sulfonation lead to three functional Brønsted acid sites: –OH, –COOH and –SO3H on the solid acid. The probe molecules experimental reveal that the acid strength of the solid acid is stronger than that of SO42− /ZrO2, but slightly weaker than that of 100% H2SO4. The catalytic performance was evaluated by the esterification of oleic acid with methanol. The results indicate that this solid acid catalyst is very active, corresponding to high conversion (72%) of esterification reaction. In addition, the spent solid acid can be recovered by simple regeneration process.  相似文献   

3.
A series of p-arsanilic acid stabilizing polyoxo‑titanium clusters (PTCs) have been successfully synthesized through solvothermal method, which present {Ti4}, {Ti5} and {Ti6} cores functionalized by different carboxylic acid ligands. PTC-111 displays a {Ti4} core structure which is stabilized by glycolic acid and esterified p-arsanilic acid. PTC-112, PTC-113 and PTC-114 have isostructural {Ti6} structures, functionalized by isonicotinic acid, 2-chloroisonicotinic acid and benzoic acid, respectively. The {Ti5} core of PTC-115 is functionalized by m-nitrobenzoic acid. It is interesting that the light absorption behaviours of these PTCs are dependent on their core structures and the functionalizing ligands.  相似文献   

4.
Due to importance and wide applications, CoCr2O4 ceramic pigment nanoparticles were synthesized via low-temperature solution combustion route by different fuels including ethylenediamine/oxalic acid, ethylenediamine/citric acid, oxalic acid/citric acid and ethylenediamine/oxalic acid/citric acid. Physicochemical properties of the synthesized samples were determined by different techniques such as fourier transform infrared spectroscopy (FT-IR), X-ray diffraction (XRD), field emission scanning electron microscopy (FE-SEM) and energy dispersive X-ray spectroscopy (EDX) and color/optical properties were evaluated based on CIELAB system by spectrophotometer. Moreover, thermodynamic considerations of combustion reactions for CoCr2O4 nanopigments formation in terms of calculated adiabatic flame temperature and enthalpy of combustion reaction were studied. The Comparison of results and data showed that cobalt chromite pigment nanoparticles synthesized by using ethylenediamine/citric acid and ethylenediamine/oxalic acid/citric acid fuels exhibited higher purity, smaller crystallite size and lower degree agglomeration.  相似文献   

5.
A series of acid modified CeO2 catalysts were prepared and used for selective catalytic reduction (SCR) of NO with NH3. The results showed that the SCR activity of pure CeO2 was greatly enhanced by the modification of acid. The CeO2 modified by 20% phosphotungstic acid exhibited the best NO conversion in a wide temperature range of 150–550 °C. The SCR activity was slightly influenced by SO2 and H2O, while such effect was reversible. The improvement of SCR activity and N2 selectivity over CeO2 catalyst modified by acid was attributed to the enhanced amount and intensity of Brønsted or Lewis acid sites.  相似文献   

6.
A rat milk substitute containing lower amounts of palmitic and oleic acid in the triacylglycerols in comparison to natural rat milk was fed to artificially reared rat pups from day 7 after birth to day 14. Pups reared by their mother served as controls. Free trideuterated (D3) palmitic acid [(C2H3)(CH2)14COOH, 98 atom % D] and free perdeuterated (D31) palmitic acid [C15 2H31COOH, 99 atom % D] in equal quantity were mixed into the triacylglycerols of the milk substitute in an amount equal to 100% of the palmitic acid in the triacylglycerols. A control milk substitute contained unlabeled free palmitic acid in an amount equal to 100% of the palmitic acid in the triacylglycerols of the milk substitute. The objective was to determine if palmitic acid in the diet contributed significantly to the palmitic acid content of developing brain and other organs. The methyl esters of the fatty acids were analyzed by gas chromatography and the palmitic acid methyl ester was examined by fast atom bombardment mass spectrometry. The proportion of deuterated methyl palmitate as a percentage of total palmitate was determined; 32% of the palmitic acid in liver and 12% of the palmitic acid in lung were trideuterated and perdeuterated palmitic acid in approximately equal amounts. The brain, by contrast, did not contain the deuterated palmitic acid moiety. Quantitation of palmitic acid and total fatty acids revealed a significant accumulation in organs in the interval from 7 to 14 days of age. Under our experimental conditions, labeled palmitic acid does not enter the brain. Consequently, we conclude that the developing brain produces all required palmitic acid byde novo synthesis.  相似文献   

7.
S. Wang  Z. Song  L. Wang 《应用陶瓷进展》2015,114(3):188-190
Single phase zirconolite (CaZrTi2O7) was synthesised by a novel method using acid corrosion to treat zirconolite–diopside mixture powders (2CaZrTi2O7–Mg2Si2O6). The mixture powders were prepared by solid state reaction method at 1200°C for 2 h. Single phase zirconolite powders were obtained by acid corrosion method with 5% hydrofluoric acid on the zirconolite–diopside Synroc mixture powders and washed with 10% acetic acid.  相似文献   

8.
Fatty acid α-hydroxylase, a cytochrome P450 enzyme, from Sphingomonas paucimobilis, utilizes various straight-chain fatty acids as substrates. We investigated whether a recombinant fatty acid α-hydroxylase is able to metabolize phytanic acid, a methyl-branched fatty acid. When phytanic acid was incubated with the recombinant enzyme in the presence of H2O2, a reaction product was detected by gas chromatography, whereas a reaction product was not detected in the absence of H2O2. When a heat-inactivated enzyme was used, a reaction product was not detected with any concentration of H2O2. Analysis of the methylated product by gas chromatography-mass spectrometry revealed a fragmentation pattern of 2-hydroxyphytanic acid methyl ester. By single-ion monitoring, the mass ion and the characteristic fragmentation ions of 2-hydroxyphytanic acid methyl ester were detected at the retention time corresponding to the time of the product observed on the gas chromatogram. The K m value for phytanic acid was approximately 50 μM, which was similar to that for myristic acid, although the calculated V max for phytanic acid was about 15-fold lower than that for myristic acid. These results indicate that a bacterial cytochrome P450 is able to oxidize phytanic acid to form 2-hydroxyphytanic acid.  相似文献   

9.
Shortenings containing ca. 1% lactic acid added as lactylated glycerides are analyzed for total or water insoluble combined lactic acid (WICLA) by a liquid-liquid partition Chromatographie pro-cedure. The sample is saponified, acidified with H2SO4, dispersed on silicic acid, slurried with chloroform and transferred to the top of a chro-matographic column composed of a 0.5 N sulfuric acid stationary phase supported on silicic acid. After elution of the fatty acids by chloroform, 15% m-butanol removes the excess H2SO4 fol-lowed by a well resolved lactic acid fraction. Titration of the latter fraction with standard methanolic sodium hydroxide has shown the lactic acid recovery to be 94%. The WICLA determina-tion requires 7 hr and has a standard deviation of ±0.020%. Presented in part at the AOCS Meeting, New Orleans, 1962.  相似文献   

10.
The effect of Ti(IV) on the degradation efficiency of acetic acid by O3/H2O2 was investigated. The removal rate of acetic acid by O3/H2O2 increased from 8.0% to 62.9% after 30 min when Ti(IV) was added to acetic acid solution at pH 2.8. The optimized parameters were as follows: the pH of acetic acid solution less than 5.0; the mass concentration ratios of H2O2 to Ti(IV), and to acetic acid about 40:1, and 3:50, respectively. The analysis of the kinetics of acetic acid degradation using the relative method showed that Ti(IV)/H2O2/O3 produced more hydroxyl radicals than did H2O2/O3.  相似文献   

11.
Lipid and fatty acid levels in the edible flesh of 17 freshwater fish from Brazil’s southern region were determined. Analyses of fatty acid methyl esters were performed by gas chromatography. Palmitic acid (C16:0) was the predominant saturated fatty acid, accounting for 50–70% of total saturated acids. Oleic acid (C18:1θ9) was the most abundant monounsaturated fatty acid. Linoleic acid (C18:2θ6), linolenic acid (C18:3θ3), and docosahexaenoic acid (C22:6θ3) were the predominant polyunsaturated fatty acids (PUFA). The data revealed that species such as truta, barbado, and corvina were good sources of eicosapentaenoic acid (C20:5θ3) and docosahexaenoic acid (C22:6θ3), and that most freshwater fish examined were good sources of PUFA θ3.  相似文献   

12.
Several Ni/SiO2 catalysts were developed for the hydrogenation of levulinic acid using formic acid as the hydrogen source. The catalysts were prepared by a variety of methods including impregnation, co‐precipitation, deposition‐precipitation, and citric acid assisted impregnation combustion. The morphological properties were investigated by XRD, N2 sorption, HRTEM, and H2 pulse chemisorption measurements. XRD patterns of the calcined material revealed the presence of NiO particles, while calcination in an inert atmosphere produced Ni particles through in situ reduction of NiO. The reaction proceeded without external H2 flow using formic acid as hydrogen source. The Ni/SiO2 catalyst prepared by the citric acid assisted method and calcined in inert gas flow was the most efficient for the hydrogenation of levulinic acid without external H2 flow. The high catalytic performance was attributed to the high dispersion of cheap and earth‐abundant Ni nanoparticles and optimal porosity.  相似文献   

13.
The Lewis acid transformation to Bronsted acid was investigated over the Pt/γ-Al2O3 hybrid catalysts in the presence of hydrogen atmosphere by in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) of adsorbed NH3. The changes of FTIR spectra were monitored during the introduction of hydrogen at 40 °C and atmospheric pressure for 130 min. The degrees of Lewis acid transformation were varied by addition of non-reducible (SiO2 and Al2O3) and reducible (ZrO2, TiO2 and CeO2) oxides to the Pt/γ-Al2O3 catalysts as the hybrid catalysts. According to the in situ DRIFTS, the hydrogen temperature programmed reduction (H2-TPR), and the hydrogen temperature programmed desorption (H2-TPD) results, the introduction of hydrogen resulted in a decrease in the amount of ammonia adsorbed on Lewis acid sites, and an increase in the amount of ammonium ions on Bronsted acid sites with time on stream. It is proposed that ammonia migration from Lewis acid sites to Bronsted acid sites occurred during the introduction of hydrogen in the presence of Pt particles when compared to the observation of only observed catalysts (without Pt particles). The addition of reducible oxides led to the high rate of Lewis acid transformation, which was higher than those of the non-reducible oxides. Weaker Lewis acid sites and higher amount of hydrogen spillover over the observed catalysts enhanced the rate of Lewis acid transformation in this study. However, the amount of Lewis acid sites at the initial stage did not play an important role in these transformations.  相似文献   

14.
The interaction of combinations of sulfur, 2,2′-dibenzothiazole (MBTS), ZnO, and stearic acid were studied by differential scanning calorimetry. A MBTS/stearic acid interaction was indicated as evidenced by the effect the MBTS/stearic acid combination had on the melting of sulfur, the Sα → Sβ transition being suppressed in favor of a Sα → Sγ transition. The dissolution/interaction of MBTS in molten sulfur was also delayed by the MBTS/stearic acid interaction, which, it is proposed, involved protonation of the N atom in MBTS by stearic acid. MBTS did not affect the formation of zinc stearate from ZnO and stearic acid, but when sulfure was added to the mixture, the ZnO/stearic acid reaction did not go to completion. No direct evidence for the formation of 2,2′-dibenzothiazole polysulphides was found, but the absence of the Sγ → Sμ transition in sulfur/MBTS mixes was interpreted as indirect evidence of a reaction between these curatives. There was no evidence for the formation of a sulfur/MBTS/ZnO compound of the type generally attributed the role of an active sulfurating agent in accelerated sulfur vulcanization.  相似文献   

15.
The effect of amount and strength of acid sites on the activity and selectivity for the selective reduction of NO with C3H6 in the presence of excess oxygen over H-form and ion-exchanged zeolites has been investigated. The activity was found to be proportional to the acid amount determined by NH3-temperature programmed desorption, but independent of the acid strength of the zeolites. The selectivity, i.e., the ratio of the reduction of NO with C3H6 to the oxidation of C3H6 by O2, was independent of the amount and strength of acid sites.  相似文献   

16.
Transition-metal Catalyzed Oxidation of Unsaturated Fatty Acids — Synthesis of Ketocarboxylic Acids and Dicarboxylic Acids Terminal unsaturated C10–C14-fatty acid methylesters (9-decenoic-, 10-un-decenoic-, 13-tetradecenoic methylesters) were converted to methylketocarboxylic methylesters (yields: 60–75%, isolated) by oxidation with O2/H2O at roomtemperature under catalysis of PdCl2/CuCl2. Using RhCl3/FeCl3 at 80°C yields of 40–60% were obtained. For the first time methyl oleate was converted directly to a mixture of 9-oxo- and 10-oxo-stearic acid methylester by palladium catalyzed oxidation. In DMF/H2O the selectivity to these two ketoesters was 85% (15% isomers), in dioxane/H2O the selectivity droped to 55% while the yield of the oxostearic acid esters climbed to 70%. The Mn-catalyzed oxidative cleavage of methylketocarboxylic acid esters with O2 at 115°C led in each case to a mixture of two dicarboxylic acid esters in a molar ratio of 2 : 1. Starting with 9-oxodecanoic acid azelaic and suberic acid were obtained at a conversion rate of 90%. Analogous 10-oxoundecanoic acid led to C10/C9- and 13-oxotetradecanoic acid led to C13/C12-dicarboxylic acids. The oxidative cleavage of 9-/10-oxostearic acid methylester yielded mixtures of C8–C10-monocarboxylic acids and methylesters of C8–C10-dicarboxylic acids.  相似文献   

17.
The effect of citric acid (CA) addition was studied on the HDS of thiophene over Co–Mo/(B)/Al2O3 catalysts. The catalysts were characterized by means of LRS, Mo K-edge EXAFS, NO adsorption capacity measurements, and UV–vis spectra. The catalysts were subjected to a chemical vapor deposition (CVD) technique using Co(CO)3NO as a precursor of Co in order to get deeper insights into the effect of citric acid addition. It was shown that the HDS activity was enhanced by the citric acid addition up to the CA/Mo mole ratio of around 1 and leveled off with further addition. The amount of Co anchored by the CVD was increased by the addition of citric acid, suggesting an increase in the dispersion of MoS2 particles on the catalyst by the simultaneous presence of Co, Mo and citric acid, in conformity with the increase in the NO adsorption capacity. In contrast to Co–Mo catalysts, the edge dispersion of MoS2 particles in Mo/B/Al2O3 was not affected by the addition of citric acid. The LRS, UV–vis spectra and Mo K-edge EXAFS showed that Co–CA and Mo–CA surface complexes are formed by the addition of citric acid. The Co–CA surface complex is more preferentially formed on CoMo/Al than on CoMo/B/Al, in agreement with a greater promoting effect of citric acid at a lower CA/Mo mole ratio for CoMo/Al than for CoMo/B/Al.  相似文献   

18.
Regeneration of spent sulphuric acid . Several processes are available for the regeneration of spent sulphuric acid. Generally, the dilute acids containing at least 20% of H2SO4 are preconcentrated to 60 – 70% and then either further concentrated to over 90% or decomposed to SO2. Purifying operations can be combined with these steps. Spent sulphuric acid highly contaminated with inorganic and organic compounds can frequently only be regenerated after a prior purification step, e. g. by extraction of electrolysis. Dilute spent sulphuric acid containing less than 20% of H2SO4 can also be regenerated, but the costs incurred by evaporation increase out of proportion with decreasing sulphuric acid concentration.  相似文献   

19.
Graft polymerization of acrylic acid (AA) onto rice starch using postassium permanganate/acid redox system as initiator was investigated. When starch was reacted with KMnO4 solution, MnO2 was deposited onto starch. The dependence of MnO2 amount deposited was directly related to KMnO4 concentration. Subjecting the MnO2-containing starch to a solution consisting of monomer (AA) and acid (citric, tartaric, oxalic and hydrochloric acid) formed poly(AA)–starch graft copolymers. The graft yield, expressed as meq COOH/100 g starch, was measured by the amount of MnO2 deposited, AA concentration, material-to-liquor ratio, kind and concentration of acid, as well as temperature and duration. Finally, the newly prepared poly(AA)–starch graft copolymers were applied to cotton textiles to determine their suitability as sizing agents. The highest graft yield was obtained with citric acid and the least with hydrochloric acid, with tartaric and oxalic acid in between. The graft yield increased by increasing the concentration of acid to a certain concentration beyond which grafting leveled off. A similar trend was observed when the magnitude of grafting was related to the amount of MnO2 deposited. The graft yield increased by increasing the polymerization temperature from 30° to 50°C. Increasing the temperature to 60°C is accompanied by decreased grafting. On the other hand, fabric samples sized with poly(AA)–starch graft copolymers acquire higher tensile strength, elongation at break, and abrasion resistance than that sized with native rice starch, i.e., poly(AA)–starch graft copolymers serve as good sizing agents for cotton textiles. A tentative mechanism for grafting rice starch with AA using the KMnO4/acid redox system was elucidated. © 1995 John Wiley & Sons, Inc.  相似文献   

20.
This paper studies the decomposition of formic, oxalic and maleic acids by O3, O3/catalyst, and O3/H2O2. The catalytic effect of Co2+, Ni2+, Cu2+, Mn2+, Zn2+, Cr3+, and Fe2+ ions is investigated. The results showed that—Co2+ and Mn2+ have the highest catalytic activity for the decomposition of oxalic acid while the catalytic effect of the studied ions is insignificant on the rate of decomposition of formic acid. Maleic acid decomposes by ozone into formic acid and glyoxylic acid, which subsequently oxidizes to oxalic acid. Though the studied ions have no effect on the decomposition of maleic acid, they have a significant effect on the produced oxalic and glyoxylic acids. In the presence of Cu2+ ions glyoxylic acid is mainly transformed into formic acid and traces of oxalic acid. In such case, a complete decomposition of maleic acid and its degradation products is achieved within 45 min. The results also show that combining H2O2 with O3 results in an increase in the rate of decomposition of oxalic acid. However, O3/H2O2 system is less active than O3/Co2+ or O3/Mn2+.  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