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 687 毫秒
1.
The CO/H2 homologation of methanol to acetaldehyde and subsequently to its dimethyl acetal in the presence of cobalt acetate promoted by iodine was examined under various conditions. Temperature and pressure were found as critical parameters. High pressures (140 MPa) and low temperatures (160–170°C) give optimal yields and selectivity to acetaldehyde. According to pressure, temperature, contact time, gas ratio and ligand/catalyst ratio, the reaction is oriented towards acetaldehyde, its dimethyl acetal or methyl acetate.  相似文献   

2.
Upon the addition of Lewis and protonic acids, promoters to ruthenium carbonyl iodide catalysts ([Ru(CO)3I3]), important effects are observed on the activity and selectivity of the hydrogenation, carbonylation and homologation reactions of methyl acetate.The improvement in selectivity to the valuable products acetic acid and ethyl acetate, and the decrease of the formation of hydrocarbons are related to the acceleration by Lewis acids of the alkyl migration-carbonyl insertion step of the process.The effect on the kinetics is related to the formation and maintenance in the catalytic solution of the hydrido species, HRu(CO)3I3, and to the availability of I involved in the activation of the substrates.  相似文献   

3.
《Tetrahedron letters》1988,29(37):4743-4746
Oxetanocin has been synthesized starting from cis-2-buten-1,4-diol through α- or β-D-oxetanosyl acetate as an important intermediate which has an α-(methyl oxalyloxy)methyl group at C2-position.  相似文献   

4.
Although industrialized, the mechanism for catalytic upgrading of bioethanol over solid‐acid catalysts (that is, the ethanol‐to‐hydrocarbons (ETH) reaction) has not yet been fully resolved. Moreover, mechanistic understanding of the ETH reaction relies heavily on its well‐known “sister‐reaction” the methanol‐to‐hydrocarbons (MTH) process. However, the MTH process possesses a C1‐entity reactant and cannot, therefore, shed any light on the homologation reaction sequence. The reaction and deactivation mechanism of the zeolite H‐ZSM‐5‐catalyzed ETH process was elucidated using a combination of complementary solid‐state NMR and operando UV/Vis diffuse reflectance spectroscopy, coupled with on‐line mass spectrometry. This approach establishes the existence of a homologation reaction sequence through analysis of the pattern of the identified reactive and deactivated species. Furthermore, and in contrast to the MTH process, the deficiency of any olefinic‐hydrocarbon pool species (that is, the olefin cycle) during the ETH process is also noted.  相似文献   

5.
The densities and speeds of sound for binary mixtures containing the solute ionic liquid (IL) methyltrioctylammonium bis(trifluoromethylsulfonyl)imide ([MOA]+[Tf2N]), solute/solvent methanol, and solvent methyl acetate have been measured at 298.15, 303.15, 308.15 and 313.15 K at atmospheric pressure. The binary mixtures studied are ([MOA]+[Tf2N] + methyl acetate or methanol), and (methanol + methyl acetate). The apparent molar volume, V φ and the apparent molar isentropic compressibility, k φ , have been evaluated from the experimental density and speed of sound data, respectively. The parameters of a Redlich–Mayer type equation were fitted to the apparent molar volume and apparent molar isentropic compressibility data. The apparent molar volume and apparent molar isentropic compressibility at infinite dilution, Vf0V_{\phi}^{0} and kf0k_{\phi}^{0}, respectively, of the binary solutions have also been calculated at each temperature. The infinite dilution apparent molar volume indicates that intermolecular interactions for (IL + methyl acetate) mixtures are stronger than for (IL + methanol) mixtures at all temperatures except at 298.15 K, and that Vf0V_{\phi}^{0} for the (IL + methyl acetate or methanol) binary systems increases with an increase in temperature. For the (methanol + methyl acetate) system the intermolecular interaction are weaker and Vf0V_{\phi}^{0} also increases with an increase in temperature. Values of the infinite dilution apparent molar expansibility, Ef0E_{\phi}^{0}, indicate that the interaction between (IL + methyl acetate) is greater than for (IL + methanol) and (methanol + methyl acetate).  相似文献   

6.
A Ni-catalyzed Csp2−OMe ortho-functionalization methodology to form chemoselectively alkyne monoannulation or aromatic homologation products is reported as a novel protocol towards the valorisation of substrates containing Csp2−OMe units. Double activation of Csp2−OMe and Csp2−F bonds is also demonstrated. Further use of aromatic homologation products towards the synthesis of nanographene-like compounds is described.  相似文献   

7.
Acetic acid (AA) has been largely used with a wide range of applications such as a raw material for a synthesis of vinyl acetate monomer, cellulose acetate or acetate anhydrate, acetate ester and a solvent for a synthesis of terephthalic acid and so on. The present paper briefly summarizes the commercialized chemical processes with their Rh or Ir-based catalytic systems in a liquid-phase carbonylation reaction such as Monsanto, Cativa and Acetica processes. In addition, some alternative catalytic systems such as heterogeneous catalysts to produce AA by direct oxidation or indirect carbonylation of dimethyl ether through BP-SaaBre process in a gas-phase reaction to solve some problems such as a difficult separation of homogeneous catalysts in a corrosive reaction medium. Some home-made heterogeneous catalysts such as a rhodium incorporated graphitic carbon nitride (Rh-g-C3N4) and some heterogenized homogeneous catalysts using the supports of tungsten carbide, iron oxide or graphitic carbon nitride containing rhodium complexes were also introduced for the synthesis of AA through a liquid-phase methanol carbonylation reaction to effectively solve the leaching problem of active rhodium metal as well as to mitigate the separation problem of homogeneous catalysts.  相似文献   

8.
Ni/KIT-6 catalysts loaded with different amounts of metallic Ni were prepared by impregnation method. The prepared catalysts and their precursors were investigated through wide- and low-angle XRD, TEM, BET, H2-TPR, and H2-TPD analyzes. The catalytic hydrodeoxygenation performance of the catalysts was evaluated using ethyl acetate as a model bio oil compound. Results indicate that the catalytic hydrodeoxygenation performance of the prepared catalysts was directly related to hydrogen storage properties, hydrogen desorption properties, dispersion of the active component Ni, and so on. The ethyl acetate conversion and ethane selectivity of 25?wt% Ni/KIT-6 catalyst were 100 and 96.8%, respectively, at 300?°C, which shows the best performance. The hydrodeoxygenation activity of ethyl acetate was higher than that of methyl acetate and isopropyl acetate because of the effect of molecular polarity and size. And, this reaction is a structure sensitive reaction.  相似文献   

9.
This work continues a series of our studies on the synthesis of nanostructured carbon (NSC) by the pyrolysis of H2 + C3–C4 alkane mixtures on nickel and cobalt metal catalysts supported on chemically diverse inorganic materials (aluminosilicates, alumina, carbon) having different textural characteristics (mesoporous and macroporous supports) and shapes (granules, foamed materials, and honeycomb monoliths). Here, we consider Ni catalysts supported on granular mesoporous silica (SiO2). It has been elucidated how the yield of synthesized carbon depends on the Ni/SiO2 catalyst preparation method (homogeneous precipitation or impregnation) and on the composition of the impregnating solution, including the molar ratio of its components—nickel nitrate and urea. The morphology of catalytic NSC and Ni distribution in the silica granule have been investigated using a scanning electron microscope with an EDX analyzer. Carbon-containing composite supports (NSC/SiO2) have been employed as adsorbents for immobilizing microbial lipase. The enzymatic activity and stability of the resulting biocatalysts have been estimated in transesterification reactions of vegetable (sunflower and linseed) oils involving methyl or ethyl acetate, esterification, and synthesis of capric acid–isoamyl alcohol esters in nonaqueous media.  相似文献   

10.
The Cu/γ-Al2O3 catalysts with different Cu loadings were prepared by impregnation method. The physicochemical properties of these Cu/γ-Al2O3 catalysts were characterized by H2-TPR, XRD, and in-situ XPS. The catalytic hydrogenation performances of methyl laurate over Cu/γ-Al2O3 catalysts were studied. The results show that the hydrogenation performances of methyl laurate on Cu/γ-Al2O3 catalyst are related to the dispersion, crystallite size, and content of the active component Cu0. The 10CA catalyst has the best hydrogenation performances for methyl laurate to produce C12 alcohol. At 300 °C, the conversion of methyl laurate and the selectivity of C12 alcohol are 55.6% and 30.4%, respectively.  相似文献   

11.
Iodide and acetate salts increase the rate of reaction of Li[RhI2(CO)2] with MeI at 25° C in acetic acid solution, a model of the rate-determining step in catalytic methanol carbonylation. The effects of water, LiBF4, and other additives are also reported. Iodide salts also promote catalytic methanol carbonylation at low water concentrations. In the case of LiI promoter, lithium acetate is produced in catalytic solutions via the reaction of LiI with methyl acetate. The promotional effects of iodide and acetate on both the model and catalytic systems are rationalized in terms of iodide or acetate coordination to [RhI2(CO)2], to yield five-coordinate RhI anions as reactive intermediates for rate-determining reactions with MeI.  相似文献   

12.
The performances of ZnO–Cr2O3+silica-alumina physically mixed and Pd impregnated on silica-alumina catalysts in the transformation of synthesis gas to hydrocarbons are compared in the present work. ZnO–Cr2O3 or Pd and silicaalumina are used as methanol synthesis and the hydrocarbon formation catalysts, respectively. The highest CO conversion corresponds to the highest relative amount of methanol synthesis active sites. The highest proximity between both types of active sites in the Pd imprenated on silica-alumina produces higher hydrocarbon selectivity and higher C1 fraction than when using the physically mixed ZnO–Cr2O3+silica-alumina catalysts.  相似文献   

13.
The syntheses of nine new 5‐iodosalicylic acid‐based 1,3,4‐oxadiazoline derivatives starting from methyl salicylate are described. These compounds are 2‐[4‐acetyl‐5‐methyl‐5‐(3‐nitrophenyl)‐4,5‐dihydro‐1,3,4‐oxadiazol‐2‐yl]‐4‐iodophenyl acetate ( 6a ), 2‐[4‐acetyl‐5‐methyl‐5‐(4‐nitrophenyl)‐4,5‐dihydro‐1,3,4‐oxadiazol‐2‐yl]‐4‐iodophenyl acetate ( 6b ), 2‐(4‐acetyl‐5‐methyl‐5‐phenyl‐4,5‐dihydro‐1,3,4‐oxadiazol‐2‐yl)‐4‐iodophenyl acetate, C19H17IN2O4 ( 6c ), 2‐[4‐acetyl‐5‐(4‐fluorophenyl)‐5‐methyl‐4,5‐dihydro‐1,3,4‐oxadiazol‐2‐yl]‐4‐iodophenyl acetate, C19H16FIN2O4 ( 6d ), 2‐[4‐acetyl‐5‐(4‐chlorophenyl)‐5‐methyl‐4,5‐dihydro‐1,3,4‐oxadiazol‐2‐yl]‐4‐iodophenyl acetate, C19H16ClIN2O4 ( 6e ), 2‐[4‐acetyl‐5‐(3‐bromophenyl)‐5‐methyl‐4,5‐dihydro‐1,3,4‐oxadiazol‐2‐yl]‐4‐iodophenyl acetate ( 6f ), 2‐[4‐acetyl‐5‐(4‐bromophenyl)‐5‐methyl‐4,5‐dihydro‐1,3,4‐oxadiazol‐2‐yl]‐4‐iodophenyl acetate ( 6g ), 2‐[4‐acetyl‐5‐methyl‐5‐(4‐methylphenyl)‐4,5‐dihydro‐1,3,4‐oxadiazol‐2‐yl]‐4‐iodophenyl acetate ( 6h ) and 2‐[5‐(4‐acetamidophenyl)‐4‐acetyl‐5‐methyl‐4,5‐dihydro‐1,3,4‐oxadiazol‐2‐yl]‐4‐iodophenyl acetate ( 6i ). The compounds were characterized by mass, 1H NMR and 13C NMR spectroscopies. Single‐crystal X‐ray diffraction studies were also carried out for 6c , 6d and 6e . Compounds 6c and 6d are isomorphous, with the 1,3,4‐oxadiazoline ring having an envelope conformation, where the disubstituted C atom is the flap. The packing is determined by C—H…O, C—H…π and I…π interactions. For 6e , the 1,3,4‐oxadiazoline ring is almost planar. In the packing, Cl…π interactions are observed, while the I atom is not involved in short interactions. Compounds 6d , 6e , 6f and 6h show good inhibiting abilities on the human cancer cell lines KB and Hep‐G2, with IC50 values of 0.9–4.5 µM.  相似文献   

14.
Abstract

We have recently shown1,2 that the Ru(II)-Sn(II) bimetallic complex can catalyze the unprecedented one-step formation of acetic acid (or methyl acetate) with methanol used as the sole source. It was suggested that the reaction consists of sequential processes of methanol → formaldehyde (methylal) → methyl formate → acetic acid (methyl acetate). While the Ru(II) complexes capable of catalyzing the dehydrogenation of methanol into methyl formate are known,3–5 this catalyst system is unique because of its extra ability to isomerize methyl formate to acetic acid without a CO atmosphere (usually high pressure) or an iodide promoter (often corrosive to reaction apparatus).6 In this communication, we examine the cyclopentadienyl bis(triphenylphosphine) ruthenium(II) auxilliary in view of its well-defined geometry and configurational stability,7 and demonstrate that combination with the SnF3 ? ligand8 gives quite high catalytic ability compared to the conventional9 SnCl3 ? ligand.  相似文献   

15.
The synthesis of dimethyl carbonate (DMC) from methanol and supercritical carbon dioxide over various base catalysts has been studied. Compounds of group-I elements (Li, Na and K) were used as base catalysts. The promoter and the dehydrating agent were also used to enhance the yield of DMC. The effects of the catalysts, promoter and dehydrating agent on the yield of DMC were investigated. By-products such as dimethyl ether (DME) and C1–C2 hydrocarbons were formed with the DMC as a main product. The yield of DMC with different alkali metal catalysts ranked in the following order: K > Na > Li. The catalysts of the metal-CO3 compounds were more effective than the metal-OH compounds in DMC synthesis. The maximum DMC yield reached up to about 12 mol% in the presence of K2CO3 (catalyst), CH3I (promoter) and 2,2-dimethoxypropane (dehydrating agent) at 130–140°C and 200 bar. The reaction mechanism of DMC synthesis from methanol and supercritical carbon dioxide was proposed.  相似文献   

16.
A new 1,3,4‐oxadiazole bridging bent organic ligand, 2,5‐bis{5‐methyl‐2‐[(4‐pyridyl)methoxy]phenyl}‐1,3,4‐oxadiazole, C28H24N4O3, L, has been used to create three novel one‐dimensional isomorphic coordination polymers, viz. catena‐poly[[[dichloridomercury(II)]‐μ‐2,5‐bis{5‐methyl‐2‐[(4‐pyridyl)methoxy]phenyl}‐1,3,4‐oxadiazole] methanol monosolvate], {[HgCl2(C28H24N4O3)]·CH3OH}n, catena‐poly[[[dibromidomercury(II)]‐μ‐2,5‐bis{5‐methyl‐2‐[(4‐pyridyl)methoxy]phenyl}‐1,3,4‐oxadiazole] methanol monosolvate], {[HgBr2(C28H24N4O3)]·CH3OH}n, and catena‐poly[[[diiodidomercury(II)]‐μ‐2,5‐bis{5‐methyl‐2‐[(4‐pyridyl)methoxy]phenyl}‐1,3,4‐oxadiazole] methanol monosolvate], {[HgI2(C28H24N4O3)]·CH3OH}n. The free L ligand itself adopts a cis conformation, with the two terminal pyridine rings and the central oxadiazole ring almost coplanar [dihedral angles = 5.994 (7) and 9.560 (6)°]. In the HgII complexes, however, one of the flexible pyridylmethyl arms of ligand L is markedly bent and helical chains are obtained. The HgII atom lies in a distorted tetrahedral geometry defined by two pyridine N‐atom donors from two L ligands and two halide ligands. The helical chains stack together via interchain π–π interactions that expand the dimensionality of the structure from one to two. The methanol solvent molecules link to the complex polymers through O—H...N and O—H...O hydrogen bonds.  相似文献   

17.
The reaction of methyl radicals with methanol at 77 K has been studied. CH3 radicals were produced by the photolysis of diphenylamine-CH3I with UV light in a methanol glass. In the methanol samples the percentage of water or the isotopic composition of methanol were changed. The reaction of CH3 radicals with methanol obeys the law C = C0exp(−Kt ) in all cases.  相似文献   

18.
Catalytic C1 chemistry based on the activation/conversion of synthesis gas (CO+H2), methane, carbon dioxide, and methanol offers great potential for the sustainable development of hydrocarbon fuels to replace oil, coal, and natural gas. Traditional thermal catalytic processes used for C1 transformations require high temperatures and pressures, thereby carrying a significant carbon footprint. In comparison, solar‐driven C1 catalysis offers a greener and more sustainable pathway for manufacturing fuels and other commodity chemicals, although conversion efficiencies are currently too low to justify industry investment. In this Review, we highlight recent advances and milestones in light‐driven C1 chemistry, including solar Fischer–Tropsch synthesis, the water‐gas‐shift reaction, CO2 hydrogenation, as well as methane and methanol conversion reactions. Particular emphasis is placed on the rational design of catalysts, structure–reactivity relationships, as well as reaction mechanisms. Strategies for scaling up solar‐driven C1 processes are also discussed.  相似文献   

19.
Fexofenadine [systematic name: (±)‐(4‐{1‐hydroxy‐4‐[4‐(hydroxydiphenyl­methyl)­piperidinium‐1‐yl]‐butyl}phenyl)‐2‐methyl­propionate], crystallizes in two forms, viz. as the methanol disolvate, C32H39NO4·2CH4O, and as the dihydrate, C32H39NO4·2H2O. It exists in the two structures as a zwitterion, which self‐assembles as dimers sustained by a pair of charged‐assisted N—H⋯OOC hydrogen bonds. In the methanol disolvate, the supramolecular organization consists of discrete fexofenadine dimers solvated by four mol­ecules of methanol. The dihydrate structure is sustained by a more extended hydrogen‐bonding scheme, wherein the hydrated dimeric entities are inter­linked by additional hydrogen bonds. The fexofenadine mol­ecule adopts different and differently disordered conformations of the 1‐hydroxy­butyl residue in the two structures.  相似文献   

20.

Oilseed crop with high oil content and promising ecological adaptability are potential sources for competitive biodiesel production. This study investigates the scope of utilizing biodiesel development through the methyl and ethyl ester from soybean and mustard oil as an alternative fuel. Methyl and ethyl esters of oils having different fatty acids compositions such as soybean (SOME and SOEE) and mustard oil (MUME and MUEE) were prepared by transesterification with methanol and ethanol in the presence of an alkali-KOH catalyst. The gas chromatographic (GC) analysis of oil samples revealed that primary fatty acid composition in soybean oil was linoleic acid (C18:2, 51.93%), followed by oleic acid (C18:1, 22.82%), palmitic acid (C16:0, 11.56%), linolenic acid (C18:3, 5.95%) and stearic acid (C18:0, 4.32%). Whereas, the main components in mustard oil were erucic acid (C22:1, 32.81%), oleic acid (C18:1, 24.98%), eicosenoic acid (C20:1, 10.44%), linolenic acid (C18:3, 8.61%) and palmitic acid (C16:0, 2.80%). The physicochemical properties (acid value, iodine value, calorific value, flash point, pour point etc.) of methyl and ethyl ester samples were estimated and found to be within the acceptable range of ASTM D6751 standards specifications. The prepared esters and oil samples were examined for cold flow properties by differential scanning calorimetry (DSC). Results revealed better cold flow properties for MUME (−2.55 °C) and MUEE (−3.10 °C) than SOME (3.21 °C) and SOEE (1.83 °C) due to more unsaturated fatty acid content in MU. Thermal and oxidative stability of samples was determined by thermogravimetric analysis (TG) and differential thermal analysis (DTA). The thermal and oxidative stability ranking of the samples was in the order of oil > methyl esters > ethyl esters.

  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