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 liquid-solid equilibrium temperature in the binary systems between n-dotriacontane and methyl or ethyl octadecanoate and methyl nonadecanoate are determined in order to obtain the interchange parameters between CH3 or CH2 and COO groups, by means of the statistics of group interaction.  相似文献   

2.
The liquid—solid equilibrium temperatures in the binary systems between n-octacosane and methyl or ethyl octadecanoate are determined in order to obtain the interchange parameters between CH3 or CH2 and COO groups by means of the statistics of group interaction.  相似文献   

3.
The phase behavior of 2,3-disubstituted methyl octadecanoate monolayers at the air-water interface is studied by film balance and a Brewster angle microscope (BAM). The comparison of the surface pressure-molecular area (pi-A) isotherms with the corresponding BAM images provides information on the phase behavior of the monolayers. Variations in the phase behavior of different 2,3-disubstituted methyl octadecanoate monolayers can be correlated with the size of the headgroups, the interactions between the polar molecular moieties and the subphase, and the intermolecular interactions. The enlarging of the headgroups makes forming a condensed monolayer difficult for the molecules, even after introduction of substituents giving rise to the formation of hydrogen bonds between the molecules, which may balance the steric repulsion and stabilize the monolayers. Model calculations of the two-dimensional lattice structure of the 2,3-disubstituted methyl octadecanoates on basis of the pg and p1 space group are performed and correspond well with the experimental results.  相似文献   

4.
The ion-molecule reactions between [CH3X]+˙ [CH3XH] +, [CH3XCH3]+ ions (X = F, Cl, Br, I) and a number of nucleophiles have been studied by ion cyclotron resonance techniques. Protonation of the nucleophiles is observed to occur from both the molecular ions [CH3]X+˙ and protonated species [CH3XH]+ whereas dimethylhalonium ions [CH3XCH3]+ react principally by methyl cation transfer. A notable exception occurs in methyl iodide where the molecular ions [CH3I]+˙ act both as proton and methyl cation donors, whereas dimethyliodonium ions are found unreactive. The results are discussed with reference to the use of alkyl halides as reagent gases in chemical ionization experiments.  相似文献   

5.
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   

6.
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   

7.
The potential energy distributions and normal coordinates (L?1 matrices) for twelve methyl halides, CH3X, CH2DX, CD2HX and CD3X (X = Cl, Br, I) have been calculated from known structural data. General harmonic force fields for methyl chloride, bromide and iodide previously determined from the most complete available isotopic frequency, Coriolis and centrifugal distortion data were used. The vibrational modes of these molecules are compared and discussed.  相似文献   

8.
Unidirectional methods are proposed for the synthesis of the 2,3-, 2,4-, and 3,4-di-, and 2,3,4-tri-O-methyl ethers of methyl (methyl α-D-galactopyranosid) uronate by the oxidation (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-galactosid) uronic acids followed by esterification with CH2N2 and catalytic hydrogenolysis of the benzyl groups.  相似文献   

9.
Deuteration measurements have been made with t-C4H9OK and with 4 moles of CH3COOH mixed with one mole of CF3COOH against 2D-selenophen, 3D-selenophen, 3CH3, 2D-selenophen and 5CH3, 2D-selenophen. Reactivity in the selenohpen-thiopen series is examined via protophilic and electrophilic isotopic-exchange reactions. The factors for the partial exchange rates with acid and base for heterocyclics with D at position 2 and CH3 at position 3, 4, or 5 are compared with those factors for isomers of monodeuterotoluene; it is found that the electronic effect of the CH3 group is transmitted similarly for these heterocyclics and for the benzene ring. Reasonably good agreement is found between the relative constants for deuterium exchange in thiophen, selenophen, and methyl derivatives of these as catalyzed by alkali-metal t-butylates in dimethylsulfoxide (DMSO) and in t-butanol mixed with diethylene glycol dimethyl ether. This shows that the results with DMSO correctly characterize the reactivity in protophilic hydrogen exchange.  相似文献   

10.
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   

11.
The thermal dissociation of the atmospheric constituent methyl formate was probed by coupling pyrolysis with imaging photoelectron photoion coincidence spectroscopy (iPEPICO) using synchrotron VUV radiation at the Swiss Light Source (SLS). iPEPICO allows threshold photoelectron spectra to be obtained for pyrolysis products, distinguishing isomers and separating ionic and neutral dissociation pathways. In this work, the pyrolysis products of dilute methyl formate, CH3OC(O)H, were elucidated to be CH3OH + CO, 2 CH2O and CH4 + CO2 as in part distinct from the dissociation of the radical cation (CH3OH+• + CO and CH2OH+ + HCO). Density functional theory, CCSD(T), and CBS-QB3 calculations were used to describe the experimentally observed reaction mechanisms, and the thermal decomposition kinetics and the competition between the reaction channels are addressed in a statistical model. One result of the theoretical model is that CH2O formation was predicted to come directly from methyl formate at temperatures below 1200 K, while above 1800 K, it is formed primarily from the thermal decomposition of methanol.  相似文献   

12.
Conclusions The telomerization of propylene with methyl dichloroacetate includes chain transfer via the C-H and C-Cl bonds, and leads to the formation of four series of telomers, the methyl esters of chloro-substituted acids of type X[CH(CH3)CH2]nCClYCO2CH3 (X, Y=H, Cl), and also of the telomeric esters of the higher alcohols and dichloroacetic acid.Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 9, pp. 2150–2152, September, 1977.  相似文献   

13.
In the absence of a direct experimental determination of A0 for a symmetric top methyl group molecule, the relations used to determine the off-axis hydrogen coordinates by the mono-substitution technique are shown to lead to the prediction of abnormally long CH bonds. The molecules methyl iodide and methane, for which A0 constants for the CH3I and CH3D species have been determined directly, and for which the ground and equilibrium state geometries are well known, are used as test cases. Methyl chloride, methyl cyanide, and methyl acetylene, for which reliable ground state geometries are known, are also considered.In order to obtain the A0 rotational constant for CH2DI required in this study, an analysis of the rovibration spectrum, of CH2DI was carried out. The results of this work are given in the Appendix.  相似文献   

14.
《Vibrational Spectroscopy》2004,34(1):169-173
The monolayer of the mixture of octadecanoic acid and octadecylamine with molar ratio 1:1 has been investigated at the air–water interface. It was found that the monolayer shows a rather stable state at the surface pressure of 30 mN/m and this monolayer can be transferred onto a CaF2 plate by Langmuir–Blodgett (LB) technique. The infrared spectra of LB films indicated that octadecylammonium octadecanoate is formed by an intermolecular proton exchange between adjacent carboxylic and aminic groups (COO and NH3+). In three-layer LB film, the CH2 scissoring mode of the long hydrocarbon chains of octadecylammonium octadecanoate shows a broad band feature at about 1468 cm−1 while this vibrational mode of three-layer LB film of the mixture (1:1) of deuterated stearic acid and octadecylamine (octadecylammonium octadecanoate-d35, C18H37NH3+C17D35COO) only shows a narrow band. The broad feature of the CH2 scissoring mode in octadecylammonium octadecanoate probably originates from the coupling between the chain of stearic acid and that of octadecylamine while this kind of coupling could be completely removed in octadecylammonium octadecanoate-d35. Another conclusion presented in this paper is that there are no couplings among the chains of fatty acid or among the chains of octadecylamine in LB films of octadecylammonium octadecanoate.  相似文献   

15.
The reactions between dichloromethylgallane and the silyl sulfides (CH3)3SiSR (R = CH3, C2H5, n-C3H7, i-C3H7, Ph and CH2Ph) in benzene result in the formation of the hitherto unknown, extremely moisture-sensitive chloro(methyl)(organylthio)gallanes. Reactions of these thiogallanes with the donor compounds N(CH3)3 and O(CH3)2 are reported. Spectra as well as some physical and chemical properties of the new compounds are given.  相似文献   

16.
Threshold photoelectron-photoion coincidence (TPEPICO) spectroscopy has been used to investigate the unimolecular chemistry of gas-phase methyl 2-methyl butanoate ions [CH3CH2CH(CH3)COOCH3·+]. This ester ion isomerizes to a lower energy distonic ion [CH2CH2CH(CH3)COHOCH3·+] prior to dissociating by the loss of C2H4. The asymmetric time of flight distributions, which arise from the slow rate of dissociation at low ion energies, provide information about the ion dissociation rates. By modeling these rates with assumed k(E) functions, the thermal energy distribution for room temperature sample, and the analyzer function for threshold electrons, it was possible to extract the dissociative photoionization threshold for methyl 2-methyl butanoate which at 0 K is 9.80 ± 0.01 eV as well as the dissociation barrier of the distonic ion of 0.86 ± 0.01 eV. By combining these with an estimated heat of formation of methyl 2-methyl butanoate, we derive a 0 K heat of formation of the distonic ion CH2CH2CH(CH3)COHOCH3·+ of 101.0 ± 2.0 kcal/mol. The product ion is the enol of methyl propionate, CH3CHCOHOCH3·+, which has a derived heat of formation at 0 K of 106.0 ± 2.0 kcal/mol.  相似文献   

17.
An investigation of the solution polymerization of methyl, butyl, isobutyl, sec-butyl, and tert-butyl methacrylates and the polymerization of methyl and butyl methacrylates in the presence of methyl, butyl, and tert-butyl isobutyrate and methyl pivalate showed that the complex order of the initiation reaction with respect to the monomer (about 2) has its cause in the ability of the ester group in the monomer and of methyl or butyl isobutyrate to activate lithium tert-alkoxide. Owing to conjugation, the ester group in the monomer is less active than the ester group in isobutyrate. Steric hindrances of the formation of a complex between lithium tert-alkoxide and ester were also investigated, because this complex is intermediate product necessary for the formation of an activated lithium tert-alkoxide, capable of initiating the polymerization of alkyl methacrylates of the type CH2?(CH3)COOCH2R.  相似文献   

18.
The crystal structure of the ethyl methyl sulfide-mercury(II) chloride complex, CH3SCH2CH3· HgCl2, has been determined by X-ray diffraction. The conformation about the CS-CC axis of the complex is trans, which is different from the conformation of crystalline ethyl methyl sulfide. The Raman and IR spectra of the complex have been measured. Observed wavenumbers of the CH2 rocking and C-C stretching vibrations of the complex are close to those of the trans form of ethyl methyl sulfide in the liquid state, but the wavenumber of the C-S stretching vibration shifts on formation of the S-Hg bond.  相似文献   

19.
The influence of the number of 3, 3, 3-trifluoropropyl(methyl)siloxane links (Φ/Φ) in the cyclotetrasiloxanes ΦmD4-m, where D represents the dimethylsiloxane link and m=0–4, on the rearrangement of these compounds in acetone solution under the action of sodium siloxanolate has been studied. The rearrangement takes place with the formation of a linear polysiloxane the degradation of which yields, in addition to the initial ring, cyclosiloxanes with a different structure. The rate of rearrangement of ΦmD4-m and of the formation of a linear polysiloxane rises with an increase in m from 0 to 3. The equilibrium concentration of the linear polysiloxane formed from ΦmD4-m is inversely proportional to m. Results have been obtained on the kinetics of the formation of the cyclosiloxanes ΦmDn, where m=0–5, n=0–5, and m+n=3–6, in the rearrangement of the rings ΦD3, Φ2D2, Φ3D, and Φ4. The reactivity of the siloxane links rises in the sequence ~ (CH3)2Si-O-Si(CH3)2 ~<~ (CF3CH2CH2)-(CH3) Si-O-Si(CH3)2 ~<(CF3CH2CH2) (CH3)Si-O-Si(CH3) (CH2CH2CF3) ~. Because of the negative inductive effect transferred through the siloxane links, the 3, 3, 3-trifluoropropyl groups strongly activate the siloxane ring with respect to nucleophiiic reagents.  相似文献   

20.
1.  The IR and Raman spectra of the methyl selenophosphonates and methyl selenophosphinates: (CH3O)2P(Se)CH3, CH3OP(Se) (C2H5)2, and CH3OP(Se) (C6H5)2 were studied in various phase states and in solvents, and with a variation in the temperature. The vibration bands of the P=Se group were assigned.
2.  A thermodynamic equilibrium of at least two steric forms of the molecule was detected in the first two compounds. The isomers arise as the result of one methoxy group rotating from the gauche position with respect to the P=Se bond to the trans position, with a gauche orientation of the other groupings.
3.  The relative amount of the isomers in the mixture and the thermodynamic parameters of the equilibrium were determined.
4.  The steric conditions created by the two aryl radicals in the third compound lead to a stabilization of one predominant conformer in all of the states of the compound, with a gauche orientation of the OCH3 with respect to the P=Se bond.
  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