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 750 毫秒
1.
We report the liquid-crystalline (LC) and luminescent properties of a series of N-annulated perylenes ( 1 – 4 ) in whose molecular structures amide and ester groups alternate. We found that the LC properties of these compounds not only depend on the number of hydrogen-bonding units, but also on the relative position of the amide linkers in the molecule. The absence of amide groups in compound 1 leads to no LC properties, whereas four amide groups induce the formation of a wide temperature range columnar hexagonal phase in compound 4 . Remarkably, compound 3 , with two amide groups in the inner part of the structure, stabilizes the columnar LC phases better than its structural isomer 2 , with the amide groups in the outer part of the molecule. Similarly, we found that only compounds 1 and 2 , which have no hydrogen bonding units in the inner part of the molecule, exhibit luminescence vapochromism upon exposure to organic solvent vapors.  相似文献   

2.
Changes in the local environment around amide groups of poly(N-isopropylacrylamide) (PNiPA) during a solvent-induced reentrant phase separation have been investigated by infrared spectroscopy combined with quantum chemical calculations. The addition of methanol or tetrahydrofuran as a cosolvent to an aqueous solution of PNiPA causes spectral changes in the amide I regions. By preparing a dimer model compound for PNiPA, we can establish the assignment of the amide I bands for the polymer in solutions. Hydrogen-deuterium exchange experiments of the amide protons of PNiPA and its dimer models have revealed that the amide groups of PNiPA form an intramolecular C=O...H-N hydrogen bond even in a good solvent. The result has suggested that the change in the amide I envelope of PNiPA observed during the solvent-induced phase transition reflects the modification of the intramolecular C=O...H-N hydrogen bond of PNiPA as well as the variation in solvation state of the amide groups. On the basis of the assignment, we have discussed contributions of the intramolecular C=O...H-N hydrogen bond to the phase behavior of PNiPA.  相似文献   

3.
Single-conformation infrared spectra in the amide I and amide II regions have been recorded for a total of 34 conformations of three α-peptides, three β-peptides, four α∕β-peptides, and one γ-peptide using resonant ion-dip infrared spectroscopy of the jet-cooled, isolated molecules. Assignments based on the amide NH stretch region were in hand, with the amide I∕II data providing additional evidence in favor of the assignments. A set of 21 conformations that represent the full range of H-bonded structures were chosen to characterize the conformational dependence of the vibrational frequencies and infrared intensities of the local amide I and amide II modes and their amide I∕I and amide II∕II coupling constants. Scaled, harmonic calculations at the DFT M05-2X∕6-31+G(d) level of theory accurately reproduce the experimental frequencies and infrared intensities in both the amide I and amide II regions. In the amide I region, Hessian reconstruction was used to extract local mode frequencies and amide I∕I coupling constants for each conformation. These local amide I frequencies are in excellent agreement with those predicted by DFT calculations on the corresponding (13)C = (18)O isotopologues. In the amide II region, potential energy distribution analysis was combined with the Hessian reconstruction scheme to extract local amide II frequencies and amide II∕II coupling constants. The agreement between these local amide II frequencies and those obtained from DFT calculations on the N-D isotopologues is slightly worse than for the corresponding comparison in the amide I region. The local mode frequencies in both regions are dictated by a combination of the direct H-bonding environment and indirect, "backside" H-bonds to the same amide group. More importantly, the sign and magnitude of the inter-amide coupling constants in both the amide I and amide II regions is shown to be characteristic of the size of the H-bonded ring linking the two amide groups. These amide I∕I and amide II∕II coupling constants remain similar in size for α-, β-, and γ-peptides despite the increasing number of C-C bonds separating the amide groups. These findings provide a simple, unifying picture for future attempts to base the calculation of both nearest-neighbor and next-nearest-neighbor coupling constants on a joint footing.  相似文献   

4.
To create a functionalized porous compound, amide group is used in porous framework to produce attractive interactions with guest molecules. To avoid hydrogen-bond formation between these amide groups our strategy was to build a three-dimensional (3D) coordination network using a tridentate amide ligand as the three-connector part. From Cd(NO3)2.4H2O and a three-connector ligand with amide groups a 3D porous coordination polymer (PCP) based on octahedral Cd(II) centers, {[Cd(4-btapa)2(NO3)2].6H2O.2DMF}n (1a), was obtained (4-btapa = 1,3,5-benzene tricarboxylic acid tris[N-(4-pyridyl)amide]). The amide groups, which act as guest interaction sites, occur on the surfaces of channels with dimensions of 4.7 x 7.3 A2. X-ray powder diffraction measurements showed that the desolvated compound (1b) selectively includes guests with a concurrent flexible structural (amorphous-to-crystalline) transformation. The highly ordered amide groups in the channels play an important role in the interaction with the guest molecules, which was confirmed by thermogravimetric analysis, adsorption/desorption measurements, and X-ray crystallography. We also performed a Knoevenagel condensation reaction catalyzed by 1a to demonstrate its selective heterogeneous base catalytic properties, which depend on the sizes of the reactants. The solid catalyst 1a maintains its crystalline framework after the reaction and is easily recycled.  相似文献   

5.
Aspartic proteinases, which include HIV‐1 proteinase, function with two aspartate carboxy groups at the active site. This relationship has been modeled in a system possessing an otherwise unactivated amide positioned between two carboxy groups. The model amide is cleaved at an enzyme‐like rate that renders the amide nonisolable at 35 °C and pH 4 owing to the joint presence of carboxy and carboxylate groups. A currently advanced theory attributing almost the entire catalytic power of enzymes to electrostatic reorganization is shown to be superfluous when suitable interatomic interactions are present. Our kinetic results are consistent with spatiotemporal concepts where embedding the amide group between two carboxylic moieties in proper geometries, at distances less than the diameter of water, leads to enzyme‐like rate enhancements. Space and time are the essence of enzyme catalysis.  相似文献   

6.
A cholesteryl-based 2-(2′-hydroxyphenyl)benzoxazole (HPB) derivative 3 linked with an amide bond was prepared through an efficient synthetic pathway. The HPB, amide, and cholesteryl groups play important roles in constructing the supramolecular gel structure. UV-vis and fluorescence spectroscopy also showed that HPB and amide groups, which provide intra- and intermolecular hydrogen bonding, respectively, also contribute the recognition of fluoride anions.  相似文献   

7.
本文合成了含酰胺基团和不含酰胺基团的两类Gemini阳离子表面活性剂,测定了其表面活性参数,研究了酰胺基团对表面活性剂的表面活性和聚集行为的影响。结果表明,酰胺基团提高了Gemini阳离子表面活性剂的临界胶团浓度,降低了胶团聚集数,增强了胶团微极性,增大了表面活性剂的饱和吸附量。  相似文献   

8.
Two liquid chromatographic chiral stationary phases based on (+)‐(18‐crown‐6)‐2,3,11,12‐tetracarboxylic acid were applied to the resolution of the amide derivatives of cyclic α‐amino acids including proline and pipecolic acid. Among the five amide derivatives of proline, aniline amide was resolved best on the first chiral stationary phase, which contains two N–H tethering amide groups, with the separation factor of 1.31 and the resolution of 2.60, and on the second chiral stationary phase, which contains two N–CH3 tethering amide groups, with the separation factor of 1.57 and the resolution of 5.50. Among the five amide derivatives of pipecolic acid, 2‐naphthyl amide was resolved best on the first chiral stationary phase with the separation factor of 1.30 and the resolution of 1.75, but 1‐naphthylmethyl amide was resolved best on the second chiral stationary phase with the separation factor of 1.30 and the resolution of 2.26. In general, the second chiral stationary phase was found to be better than the first chiral stationary phase in the resolution of the amide derivatives of cyclic α‐amino acids. In this study, the second chiral stationary phase was first demonstrated to be useful for the resolution of secondary amino compounds.  相似文献   

9.
In the last thirty years, Gemini surfactants with various structures have been designed, synthesized, and demonstrated to show superior physicochemical properties. However, the utilization of non-degradable surfactants, including these Gemini surfactants, poses a threat to the environment; hence, degradable Gemini surfactants are desirable. Herein, biodegradable cationic Gemini surfactants with amide or ester groups in the hydrophobic chains or the spacer were synthesized. A monomeric surfactant containing an amide group and a Gemini surfactant with amide groups both in the hydrophobic chains and the spacer were synthesized for comparison. The effects of amide group location on the aggregation behavior of Gemini surfactants were studied systematically. The differences between the Gemini surfactants with amide groups and Gemini surfactants with ester groups were evaluated by comparing their aggregation behavior and hydrogen bonding formation. The Gemini surfactants with amide groups (C12A-Cn-AC12) in the chains showed much larger exothermic ΔHmic and more negative ΔGmic values than those of the corresponding monomeric surfactant C12A; besides, their critical micelle concentration (cmc) was more than one order of magnitude lower than that of C12A. The amide groups located in the hydrophobic alkyl chains promoted hydrogen bonding formation and self-assembly of the Gemini surfactants C12A-Cn-AC12. Moreover, 1H NMR spectra revealed that the co-effect of a short spacer and hydrogen bonding leads to slow exchange of the C12A-C2-AC12 molecules between the monomer and the aggregate. For the Gemini surfactant series C12-ACnA-C12, the amide groups notably increased the spacer length, and largest cmc value and smallest exothermic ΔHmic value were observed for C12-AC2A-C12 instead of C12-AC6A-C12. In C12-AC12A-C12, the spacer was long and sufficiently flexible to adopt a "U"-shaped conformation above the cmc, and it acted as the hydrophobic part of the surfactant, as confirmed by 1H NMR spectra. Among the Gemini surfactant with amide groups in both the spacer and the hydrophobic alkyl chains, C12A-AC6A-AC12 had a smaller cmc and I1/I3 ratio as well as more exothermic ΔHmic values than those of C12A-C6-AC12 and C12-AC6A-C12. 1H NMR spectra indicated that an ester-alcohol structural equilibrium exists during aggregation for the Gemini surfactants with ester groups. In addition, the Gemini surfactants with ester groups formed water-mediated hydrogen bonds in the aggregates. This water-mediated hydrogen bonding between ester groups was weaker than the direct hydrogen bonding between amide groups. Therefore, the Gemini surfactants with ester groups, C12E-C6-EC12 and C12-EC6E-C12, exhibited lower surface activity, a larger micelle ionization degree, higher micropolarity, and smaller exothermic ΔHmic and less negative ΔGmic values than their counterparts with amide groups, C12A-C6-AC12 and C12-AC6A-C12.  相似文献   

10.
The wetting behavior of a series of aliphatic polyamides (PAs) has been examined. PAs with varying amide content and polyethylene (PE) were molded against glass to produce surfaces with similar roughness. After cleaning, chemical composition of the surfaces was verified with X-ray photoelectron spectroscopy. Advancing and receding contact angles were measured from small sessile water drops. Contact angles decreased with amide content while hysteresis increased. Hysteresis arose primarily from molecular interactions between the contact liquid and the solid substrates, rather than moisture absorption, variations in crystallinity, surface deformation, roughness, reorientation of amide groups, or surface contamination. Free energies of hysteresis were calculated from contact angles. For PE, which is composed entirely of nonpolar methylene groups, free energies were equivalent to the strength of dispersive van der Waals bonds. For PAs, free energies corresponded to fractional contributions from the dispersive methylene groups and polar amide groups.  相似文献   

11.
Reactions of N-methyl-triorganyliminophosphoranes in Liquid Ammonia The reaction of N-methyl-triorganyliminiophosphoranes with potassium amide in liquid ammonia results in nucleophilic displacement of organyl groups by amide ions or in elimination of alkens, depending on type of organyl groups bonded to phosphorus. The reactions are compared with those of corresponding compounds nonsubstituted on nitrogen.  相似文献   

12.
A series of novel tolane‐type liquid crystals with amide group as terminal group have been prepared. The terminal amide groups were modified, and the influence of these structural parameters on liquid crystal phases was investigated by polarizing optical microscope (POM) and differential scanning calorimetry (DSC). Three of these new compounds exhibit nematic phase, good thermal stabilities. In general, these liquid crystals with amide as end groups have high melting points and phase transition temperatures, which result from the hydrogen bonds. Based on theoretical calculations, these new molecules with strong electron donating amide as end group have narrower HOMO‐LUMO energy gap and higher dipole moment than tolane.  相似文献   

13.
The title compound, C9H15N3O3, which has crystallographically imposed threefold symmetry, crystallizes as a hexagonal columnar structure. The crystal structure is stabilized by a less common amide–amide synthon, where one amide group is hydrogen bonded to four others. The amide groups form cyclic amide–amide hexamers via N—H...O hydrogen bonds.  相似文献   

14.
The reaction of polyacrylamide with formaldehyde was studied in a neutral aqueous medium at equal initial molar concentrations of amide groups and of formaldehyde (0.05 mol/L) and in a range of temperatures from 45 to 75°C. The process was investigated by measuring the loss of free formaldehyde in the reaction mixture and the changes of the sum of free formaldehyde and methylol groups versus time. The addition of HCHO to an amide function of the polymer leads to its N-methylol derivative which may transform into the product of condensation between the latter and another amide group. Because of high dilution of polyacrylamide macromolecules in the reaction mixtures studied, cross-linking of the polymer chains with formaldehyde is rather unlikely. Therefore the disappearance of the N-methylols formed is probably due to some intramolecular reactions. It is believed that they involve the condensation of N-hydroxymethyls with neighboring amide groups which results in cyclic structures containing methylenediamide sequences. The occurrence of intramolecular reactions was confirmed by applying Flory's theory of gelation. The addition of HCHO to amide functions is a rate-determining stage in the case of polyacrylamide. For this reaction the rate constants were estimated and the corresponding activation energy was found to be 62 kJ/mol.  相似文献   

15.
The conditions of preparation of mixed cellulose ethers containing carboxymethyl and amide groups by reaction with urea were examined with samples of commercial carboxymethyl cellulose and that prepared from flax cellulose as examples. The water solubility of the resulting products and their sorption activity toward copper ions were examined in relation to the amount of amide groups introduced.  相似文献   

16.
Values of the area per surfactant molecule of various single chain and gemini quaternary ammonium surfactants containing biodegradable amide and ester groups are obtained from the surface tension measurements and they are mutually compared. It was found that surfactant molecules with the ester group in their structure occupy smaller area at the air/water interface than the corresponding molecules with the amide group, mainly due to the higher conformational flexibility of ester groups. In decreasing the area per surfactant molecule value, hydrogen bonding (both inter- and intramolecular) plays a significant role when amide groups are present in the spacer of a gemini molecule. They must be separated by a polymethylene chain or a flexible group such as cyclohexane which is short enough to allow intramolecular hydrogen bonds. The flexible cyclohexane group with the amide group in single chain surfactants may lead to the formation of intermolecular hydrogen bonds among surfactant molecules which also results in the reduction of the area per surfactant molecule.  相似文献   

17.
Matrix-assisted laser desorption/ionization time-of-flight/time-of-flight tandem mass spectrometry (MALDI-TOF/TOF-MS/MS) was employed to analyze a poly(ester amide) sample (PEA-Bu) from the melt condensation of sebacic acid and 4-amino-1-butanol. In particular, we investigated the fragmentation pathways, the ester/amide bond sequences and the structure of species derived from side reactions during the synthesis. MALDI-TOF/TOF-MS/MS analysis was performed on cyclic species and linear oligomers terminated by dicarboxyl groups, carboxyl and hydroxyl groups and diamino alcohol groups. The sodium adducts of these oligomers were selected as precursor ions. Different end groups do not influence the fragmentation of sodiated poly(ester amide) oligomers and similar series of product ions were observed in the MALDI-TOF/TOF-MS/MS spectra. According to the structures of the most abundant product ions identified, the main cleavages proceed through a beta-hydrogen-transfer rearrangement, leading to the selective scission of the --O--CH2-- bonds. Abundant product ions originating from --CH2--CH2-- (beta-gamma) bond cleavage in the sebacate moiety were also detected. Their formation should be promoted by the presence of an alpha,beta-unsaturated ester or amide end group. MALDI-TOF/TOF-MS/MS provided structural information concerning the ester/amide sequences in the polymer chains. In the MALDI-TOF/TOF-MS/MS spectra acquired, using argon as the collision gas, of cyclic species and linear oligomers terminated by diamino alcohol groups, product ions in the low-mass range, undetected in the mass spectra acquired using air as the collision gas, proved to be diagnostic and made it possible to establish the presence of random sequences of ester and amide bonds in the poly(ester amide) sample. Furthermore, MALDI-TOF/TOF-MS/MS provided useful information to clarify the structures of precursor ions derived from side reactions during the synthesis.  相似文献   

18.
液晶性芳香酰胺化合物的合成   总被引:7,自引:0,他引:7  
合成了一系列炖粹以酰胺基为中心桥键的刚性芳香酰胺小分子化合物,并对其作了表征,发现其中有些化合物具有液晶性。酰胺键之间能形成很强的分子间氢键,使芳香酰胺小分子化合物的熔点很高,难于形成液成液晶态。研究发现,如果在这类化合物的中心苯环上引入合适的取代基以减弱分子间氢键,同时引入合适的末端基时,则可使芳香酰胺化合物生成液晶相的能力增强。  相似文献   

19.
Zhou XT  Rehman A  Li C  Savage PB 《Organic letters》2000,2(19):3015-3018
We have prepared a triamine derivative of cholic acid with protecting groups on the amines that allow sequential amide formation. The triamine was formed from 3 alpha,7 alpha, 12 alpha-trihydroxycholan-24-ol with good stereoselectivity. Sequential removal of the amine protecting groups and amide formation was achieved in high-yielding steps and was performed in solution and on a solid support.  相似文献   

20.
The title compounds contain groups (amine, amide, imine, carboxylic acid) that are capable of forming intramolecular hydrogen bonds involving a six-membered ring. In compounds where the two interacting functional groups are imine and carboxylic acid, the imine is protonated to give a zwitterion; where the two groups are imine and amide, the amide remains intact and forms a hydrogen bond to the imine nitrogen. The former is confirmed by the iminium 15N signal, which shows the coupling of 1J(15N,1H) -85 to -86.8 Hz and 3J(1H,1H) 3.7-4.2 Hz between the iminium proton and the methine proton of a cyclopropyl substituent on the iminium nitrogen. Hydrogen bonding of the amide is confirmed by its high 1H chemical shift and by coupling of the amide hydrogen to (amide) nitrogen [(1J(15N,1H) -84.7 to -90.7 Hz)] and to ortho carbons of a phenyl substituent. Data obtained from N,N-dimethylanthranilic acid show 15N-1H coupling of (-)8.2 Hz at 223 K (increasing to (-)5.3 Hz at 243 K) consistent with the presence of a N... H-O hydrogen 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