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.
The H2 and CH4 chemical ionization mass spectra of the olefinic esters methyl acrylate, methyl methacrylate, methyl crotonate, methyl 3-butenoate, methyl 2-methyl-2-butenoate, methyl 3-methyl-2-butenoate and methyl cinnamate have been determined. In addition to the expected loss of CH3OH from [MH]+, in many cases the protonated molecules also show loss of CO or CH2CO with methoxy group migration to the positive ion centre, indicative of protonation at the double bond. These rearrangement reactions, which have analogies in electron impact mass spectra, result in chemical ionization mass spectra of isomeric molecules which show more substantial differences than the electron impact mass spectra. In the case of methyl cinnamate, isotopic labelling experiments show considerable interchange of the added proton with the ortho and meta phenyl hydrogens prior to CH3OH or CH2CO loss, although the extent of interchange is not the same for both cases.  相似文献   

2.
The title compound, C6H13NO5, adopts a zwitterionic form where the carboxylic acid H atom is transferred to the amino group. The methyl–glycine backbone is planar. The tris(hydroxy­methyl)­methyl group is rotated as a rigid group around the amino–methyl bond by 22 (1)° and the carboxylic acid plane is rotated by 19.76 (12)° from the plane of the main skeleton. Apart from their H atoms, the three hydroxy­methyl groups adopt a propeller‐like conformation around the amino–methyl bond, close to C3 symmetry.  相似文献   

3.
A Cu/Cr2O3 catalyst was prepared by co-precipitation method, studied in methanol dehydrocoupling to methyl formate in different gas streams and characterized by BET, XRD, TPR, TPD of NH3 and CO2, etc. The results demonstrate that the catalyst can catalyze the dehydrocoupling of methanol to methyl formate in high efficiency,e. g. 99% selectivity to methyl formate at 48% conversion of methanol. The results further indicate that metallic copper might be the active species for the formation of methyl formate  相似文献   

4.
To better understand the chemistry of biodiesel surrogates, the gas‐phase oxidation of a C12 unsaturated methyl ester, methyl‐10‐undecenoate, has been studied in a jet‐stirred reactor in the temperature range 500–1100 K. These experiments were performed using neat fuel synthesized in the laboratory, with an initial fuel mole fraction set as 0.0021, at quasi‐atmospheric pressure (1.07 bar), at a residence time of 1.5 s with dilute mixtures in helium of equivalence ratios of 0.5, 1.0, and 2.0. The maximum obtained conversion was shown to be more than twice lower than that of methyl decanoate under the same conditions. This difference cannot be reproduced by the only published model for an unsaturated ester with a close number of carbon atoms (methyl‐9‐decenoate). A large range of products was quantified in addition to common oxidation products: saturated and unsaturated aldehydes, saturated and unsaturated methyl esters with a second carbonyl function, C2–C10 alkenes, C4–C10 dienes, C4–C10 unsaturated methyl esters, C8–C9 saturated methyl esters, and saturated, unsaturated, and hydroxyl methyl esters involving a cyclic ether. Pathways of formation for the products specific to unsaturated ester oxidation were proposed, and possible model improvements were discussed.  相似文献   

5.
The molecules of methyl 3‐(2‐nitrophenylhydrazono)butanoate, C11H13N3O4, (I), and methyl 3‐(2,4‐dinitrophenylhydrazono)butanoate, C11H12N4O6, (II), both prepared from methyl 3‐oxobutanoate and the corresponding nitrophenylhydrazine, exhibit polarized molecular electronic structures; in each of (I) and (II), the molecules are linked into chains by a single C—H...O hydrogen bond. The molecules of 5‐hydroxy‐3‐methyl‐1‐phenyl‐1H‐pyrazole, C10H10N2O, (III), prepared by the reaction of methyl 3‐oxobutanoate and phenylhydrazine, are linked into chains by a single O—H...N hydrogen bond. The reaction between methyl 3‐oxobutanoate and 3‐nitrophenylhydrazine yields 5‐hydroxy‐3‐methyl‐1‐(3‐nitrophenyl)‐1H‐pyrazole, (IV), which when crystallized from acetone yields 4‐isopropylidene‐3‐methyl‐1‐(3‐nitrophenyl)‐1H‐pyrazol‐5(4H)‐one, C13H13N3O3, (V).  相似文献   

6.
The stereoregularity of poly(methyl acrylate) and poly(methyl acrylate-αd) was determined from the NMR spectra. A method of quantitative determination of stereoregularity of poly(methyl acrylate) proposed in this paper is based on the fact that in the 100 Mc./sec. NMR spectrum the absorption peaks due to methylene protons in syndiotactic configurations overlap absorptions due to only one of two methylene protons in isotactic configurations. The stereostructure of poly(methy1 acrylates) polymerized with anionic catalysts such as Grignard reagents, n-butyllithium, and LiAlH4 is generally richer in isotactic diads than in syndiotactic diads. For example, poly(methyl acrylate) polymerized with phenylmagnesium bromide as catalyst at ?20°C. consists of 99% isotactic and 1% syndiotactic diads. In radical polymerization, the isotacticity of poly(methyl acrylate) is independent of polymerization temperature. Poly(methyl acrylates) polymerized with a Ziegler-Natta catalyst consisting of Al(C2H5)2Cl and VCl4 have configurations similar to those polymerized by radical initiators. The stereoregularity of poly(methyl acrylate-α-d) resembled that of poly(methyl acrylate) polymerized under the same conditions.  相似文献   

7.
RuCl3 catalyst system has many advantages for the hydro‐esterification of methyl formate and ethylene to methyl propionate. However, the unsatisfied performance restricts the development of this route. In this work, high‐performance RuCl3 catalyst systems (RuCl3‐[PPN]Cl‐Et4NI and RuCl3‐NaI) are firstly reported for this reaction. In RuCl3‐[PPN]Cl‐Et4NI catalyst system, the conversion of methyl formate and the selectivity to methyl propionate are 93.9% and 90.9% at mild reaction conditions (165°C, 2.5 MPa), respectively. Noticeably, a simple inorganic RuCl3‐NaI catalyst system achieves 88.8% conversion of methyl formate and 97.6% selectivity to methyl propionate (86.7% yield) at same conditions. NaI, as a promoter, may inhibit the decomposition of methyl formate and be conducive to the formation of methyl propionate. The effects of solvents and promoters are investigated in detail. In addition, the reaction mechanism has been also analyzed. It is hoped to lay a certain foundation for further industrial application.  相似文献   

8.
Mechanical treatment of the K2PtCl4 solid salt in a vibrating mill results in Pt–Cl bond heterolysis to form coordinatively unsaturated Pt(II) complexes. At room temperature, the freshly treated K2PtCl4 salt absorbs methyl bromide and evolves methyl chloride to the gas phase. The reaction mechanism involves the following sequence of steps: the oxidative addition of methyl iodide to Pt(II) with the intermediate formation of Pt(IV) methyl complexes and the decomposition of the latter due to intramolecular reductive elimination with methyl chloride formation. The first step of the reaction of MeI with the preactivated surface of the K2PtCl4 salt is assisted by active sites, which are regenerated in each act of the chemical transformation of MeI into MeCl involving in the chain substitution of halogen in methyl iodide. The coordinatively unsaturated surface platinum complexes can act as such active sites. Due to their effective positive charge, they can provide electrophilic assistance to nucleophilic substitution. Chain termination is probably due to the coordination of the complex with a coordination vacancy and an interstitial chloride ion to the inactive K2PtCl4 complex.  相似文献   

9.
It has been found that immobilized cells ofC. pasteurianum possessing hydrogenase activity efficiently catalyze reversible reduction of concentrated (up to 0.5M) solutions of methyl viologen with H2. A 0.5M aqueous solution of methyl viologen dissolves 240 times as much H2 as pure water under the same pressure of hydrogen. The experimentally obtained levels of methyl viologen reduction and H2 evolution are in satisfactory agreement with theoretical calculations. The potential of concentrated solutions of methyl viologen containing immobilized hydrogenase as a H2 storage medium is discussed.  相似文献   

10.
Chain transfer constants were obtained for styrene, methyl methacrylate, methyl acrylate and vinyl acetate, polymerized in methyl oleate and methyl stearate at 60°C. Transfer constants increased in the order: methyl methacrylate < styrene < methyl acrylate ? vinyl acetate in both solvents. Average values of the transfer parameters were: for methyl oleate, Qtr = 2.04 × 10?4, etr = 1.08; for methyl stearate, Qtr = 0.373 × 10?4, etr = 1.01. Indication that polar species predominate in the transition state is supported by the observed order of reactivity. The usual rate dependence appeared to be followed by all of the monomers except vinyl acetate, which was retarded, severely in methyl oleate. Transfer in methyl oleate was about 5.8 times greater than that found in methyl stearate for these four monomers. The internal allylic double bond of methyl oleate had about the same reactivity in transfer as had the terminal unsaturation in N-allylstearamide at 90°C. Rough estimates were obtained of the monomer transfer constants for the long side-chain homologs of these four monomers from the respective monomer transfer constants and the experimental transfer constants, corrected for transfer to the labile groups of the solvent. It was concluded that the rate of polymerization would determine in large measure the degree of polymerization for the reactive 18-carbon homologs but that the molecular weight of poly(vinyl stearate) and (oleate) will be regulated primarily by transfer to monomer.  相似文献   

11.
Quantum chemical calculations at the second‐order Moeller–Plesset (MP2) level with 6‐311++G(d,p) basis set have been performed on the lithium‐bonded and hydrogen‐bonded systems. The interaction energy, binding distance, bond length, and stretch frequency in these systems have been analyzed to study the nonadditivity of methyl group in the lithium bonding and hydrogen bonding. In the complexes involving with NH3, the introduction of one methyl group into NH3 molecule results in an increase of the strength of lithium bonding and hydrogen bonding. The insertion of two methyl groups into NH3 molecule also leads to an increase of the hydrogen bonding strength but a decrease of the lithium bonding strength relative to that of the first methyl group. The addition of three methyl groups into NH3 molecule causes the strongest hydrogen bonding and the weakest lithium bonding. Although the presence of methyl group has a different influence on the lithium bonding and hydrogen bonding, a negative nonadditivity of methyl group is found in both interactions. The effect of methyl group on the lithium bonding and hydrogen bonding has also been investigated with the natural bond orbital and atoms in molecule analyses. © 2008 Wiley Periodicals, Inc. Int J Quantum Chem, 2009  相似文献   

12.
The liquid—solid equilibrium temperatures in the binary systems between hexadecane and methyl nonadecanoate, methyl octadecanoate, ethyl octadecanoate, and methyl hexadecanoate are determined in order to obtain the interchange parameters between CH3 or CH2 and COO groups by means of group interaction statistics.  相似文献   

13.
The nonadditivity of methyl group in the single‐electron hydrogen bond of the methyl radical‐water complex has been studied with quantum chemical calculations at the UMP2/6‐311++G(2df,2p) level. The bond lengths and interaction energies have been calculated in the four complexes: CH3? H2O, CH3CH2? H2O, (CH3)2CH? H2O, and (CH3)3C? H2O. With regard to the radicals, tert‐butyl radical forms the strongest hydrogen bond, followed by iso‐propyl radical and then ethyl radical; methyl radical forms the weakest hydrogen bond. These properties exhibit an indication of nonadditivity of the methyl group in the single‐electron hydrogen bond. The degree of nonadditivity of the methyl group is generally proportional to the number of methyl group in the radical. The shortening of the C···H distance and increase of the binding energy in the (CH3)2CH? H2O and (CH3)3C? H2O complexes are less two and three times as much as those in the CH3CH2? H2O complex, respectively. The result suggests that the nonadditivity among methyl groups is negative. Natural bond orbital (NBO) and atom in molecules (AIM) analyses also support such conclusions. © 2008 Wiley Periodicals, Inc. Int J Quantum Chem, 2009  相似文献   

14.
Methyl n-alkyl ketones form a class of molecules with interesting internal dynamics in the gas-phase. They contain two methyl groups undergoing internal rotations, the acetyl methyl group and the methyl group at the end of the alkyl chain. The torsional barrier of the acetyl methyl group is of special importance, since it allows for the discrimination of the conformational structures. As part of the series, the microwave spectrum of octan-2-one was recorded in the frequency range from 2 to 40 GHz, revealing two conformers, one with C1 and one with Cs symmetry. The barriers to internal rotation of the acetyl methyl group were determined to be 233.340(28) cm−1 and 185.3490(81) cm−1, respectively, confirming the link between conformations and barrier heights already established for other methyl alkyl ketones. Extensive comparisons to molecules in the literature were carried out, and a small overview of general trends and rules concerning the acetyl methyl torsion is given. For the hexyl methyl group, the barrier height is 973.17(60) cm−1 for the C1 conformer and 979.62(69) cm−1 for the Cs conformer.  相似文献   

15.
A new catalytic reaction of the competing phenylation and hydrophenylation in air of methyl acrylate with tetraphenylantimony chloride in the presence of PdCl2 (0.04 mol per 1 mol of organometallic compound) in acetonitrile at 50°C for 6 h was studied. The yields of methyl cynnamate and methyl hydrocynnamate were 0.73 and 0.27 mol mol?1 respectively. The products ratio obtained depends slightly on the process duration, the Ph4SbCl and methyl acrylate ratio, and the structure of Pd salt [PdCl2, Pd(OAc)2, Li2PdCl4], but significantly on the nature of a solvent (MeCN > DMF > THF). The use of Ph4SbCl instead of Ph4SbBr leads to decrease in the yield of methyl hydrocynnamate to 0.04 mol mol?1. In the reactions of Ph4SbX (X = F, I, OAc, O2CEt) the product is not formed at all.  相似文献   

16.
Alternative unidirectional methods for synthesizing methyl (methyl α-D-galactopyranosid)uronate and its mono-O-methyl ethers by the oxidation (with CrO3-H2SO4-acetone) of the corresponding methyl O-benzyl-O-methyl-α-D-galactopyranosides having unsubstituted 6-OH groups to the corresponding methyl O-benzyl-O-methyl-α-D-galactouronic acids followed by esterification with CH2N2 and the catalytic hydrogenolysis of the benzyl groups are proposed.  相似文献   

17.
The comparative studies on the miscibility and phase behavior between the blends of linear and star‐shaped poly(2‐methyl‐2‐oxazoline) with poly(vinylidene fluoride) (PVDF) were carried out in this work. The linear poly(2‐methyl‐2‐oxazoline) was synthesized by the ring opening polymerization of 2‐methyl‐2‐oxazoline in the presence of methyl p‐toluenesulfonate (MeOTs) whereas the star‐shaped poly(2‐methyl‐2‐oxazoline) was synthesized with octa(3‐iodopropyl) polyhedral oligomeric silsesquioxane [(IC3H6)8Si8O12, OipPOSS] as an octafunctional initiator. The polymers with different topological structures were characterized by means of Fourier transform infrared spectroscopy and nuclear magnetic resonance spectroscopy. It is found that the star‐shaped poly(2‐methyl‐2‐oxazoline) was miscible with poly(vinylidene fluoride) (PVDF), which was evidenced by single glass‐transition temperature behavior and the equilibrium melting‐point depression. Nonetheless, the blends of linear poly(2‐methyl‐2‐oxazoline) with PVDF were phase‐separated. The difference in miscibility was ascribed to the topological effect of PMOx macromolecules on the miscibility. © 2006 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 44: 942–952, 2006  相似文献   

18.
Abstract

15-Keto-13, 14-trans-prostaglandin B1 methyl ester, 13,14-trans-prostaglandin B1 methyl ester, 13,14-cis-prostaglandin B1 methyl ester, 13,14-dihydro-prostaglandin B1 methyl ester and 13,14-dehydro-prostaglandin B1 are organic intermediates used in the synthesis of prostaglandin Bx, a polymeric derivative of 15-keto-prostaglandin B1 methyl ester. PGBx has been shown to protect laboratory animals against cardiogenic shock, cerebral ischemia and hypoxia. A normal phase, high performance liquid chromatographic analysis is presented which permits the identification and quantitation of these PGB1 intermediates.  相似文献   

19.
The Group XIV tetratolyl series X(C6H4‐CH3)4 (X=C, Si, Ge, Sn, Pb) were studied by using inelastic neutron scattering to measure the low‐energy phonon spectra to directly access the methyl‐group torsional modes. The effect of increased molecular radius as a function of the size of the central atom was shown to have direct influence on the methyl dynamics, reinforced with the findings of molecular dynamics and contact surface calculations, based upon the solid‐state structures. The torsional modes in the lightest analogue were found to be predominantly intramolecular: the Si and Ge analogues have a high degree of intermolecular methyl–methyl group interactions, whilst the heaviest analogues (Sn and Pb) showed pronounced intermolecular methyl interactions with the whole phonon bath of the lattice modes.  相似文献   

20.
Oxidative addition of methyl iodide to Vaska’s complex in the ionic liquids 1-butyl-3-methylimidazolium triflate [C4mim][OTf], [C4mim] bis(trifluormethylsulfonyl)imide [Tf2N], and N-hexylpyridinium [C6pyr][Tf2N] occurred cleanly to give the expected Ir(III) oxidative addition product. Pseudo-first order rate constants were determined for the oxidative addition reaction in each solvent ([Vaska’s] = 0.25 mM, [CH3I] = 37.5 mM). The observed rate constants under these conditions were 5-10 times slower than the rate seen in DMF. At high methyl iodide concentrations (>23 mM), the expected first order dependence on methyl iodide was not observed. In each ionic liquid, there was no change in the reaction rates within experimental error over the methyl iodide concentration range of 23-75 mM. At lower methyl iodide concentration, a decrease in rate was observed in [C4mim][Tf2N] with decreasing concentration of methyl iodide.  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