共查询到20条相似文献,搜索用时 15 毫秒
1.
Tsu‐Shang Leu Chun‐Shan Wang 《Journal of polymer science. Part A, Polymer chemistry》2001,39(23):4139-4151
The kinetics and mechanisms of the solution imidization of polyamic acid resulting from a diamine, bis(4‐aminophenoxy‐3,5‐dimethylphenyl)naphthylmethane, and a dianhydride, 3,3′4,4′‐diphenylsulfonetetracarboxylic dianhydride, were studied at three various temperatures (145, 165, and 180 °C). The results were confirmed by means of 1H NMR and gel permeation chromatography (GPC). Kinetic parameters were obtained by an isothermal study, and the results were quite close to second‐order kinetics for the homogeneous solution imidization. In addition, Carother's equation, Mark–Houwink theory, and GPC were used to explain the molecular weight of the imidization processes. The apparent activation energy (Ea) was 104 KJ/mol, and the pre‐exponential factor (k0) was 3.48 × 1014. The proposed kinetic mechanism is in good agreement with the kinetic models. © 2001 John Wiley & Sons, Inc. J Polym Sci Part A: Polym Chem 39: 4139–4151, 2001 相似文献
2.
2H NMR spin relaxation study of the soft segment motion in alternating ethylene–propylene copolymers
《Magnetic resonance in chemistry : MRC》2002,40(2):161-164
We synthesized three partially deuterated polymer samples, namely a poly(ethylene‐alt‐propylene) (EP) alternating copolymer, a poly(styrene‐b‐EP) diblock copolymer (SEP) and a poly(styrene‐b‐EP‐b‐styrene) triblock copolymer (SEPS). The 2H spin–lattice relaxation time, T1, of EP soft segments above their glass transition temperature was measured by solid‐state 2H NMR spectroscopy. It was found that the block copolymers had a fast and a slow T1 component whereas EP copolymer had only a fast component. The fast T1 components for SEP and SEPS are similar to the T1 value of EP above ca 20°C. The slow T1 component for SEP and SEPS exhibited a minimum at 60°C and approached the value of the fast component near the Tg of polystyrene. The motional behavior of the EP units for SEP is similar to that of SEPS over the entire range of temperature. Copyright © 2001 John Wiley & Sons, Ltd. 相似文献
3.
Marcus Giotto David Azar Jaimie Gosselin Paul T. Inglefield Alan A. Jones 《Journal of Polymer Science.Polymer Physics》2001,39(13):1548-1552
Carbon‐13 spin–lattice relaxation times are measured for poly(octadecyl acrylate) above and below the melting point of the crystalline side chains. The chain backbone has long spin–lattice relaxation times below the melting point that shorten by more than an order of magnitude as the melting point range is traversed. Below the melting point, the backbone is nearly immobilized with spin–lattice relaxation changing very slowly with temperature. Above the melting point, the shorter spin–lattice relaxation times are typical of a rubber above the glass transition and decrease with increasing temperature. The methylene groups in the side chain are quite mobile well below the melting point, indicating fairly rapid anisotropic motion within the crystal. The methyl group at the end of the chain and the adjacent methylene group have longer spin–lattice relaxation times, indicating the greatest side‐chain mobility at the end, which is in the middle of the crystal structure. The side‐chain carbon adjacent to the carbonyl group is as mobile as the majority of the side‐chain carbon, indicating side‐chain mobility extends to all of the side‐chain CH2 groups. The abrupt transition in the mobility of the backbone is not typical of the amorphous phase in a semicrystalline polymer where the backbone units can crystallize. The close proximity of every backbone segment to the crystalline domain locks backbone segmental motion below the melting point. Melting and crystallization of the side chains switch segmental motion of the backbone on and off. © 2001 John Wiley & Sons, Inc. J Polym Sci Part B: Polym Phys 39: 1548–1552, 2001 相似文献
4.
Meta- and para-diethyl-p,p-oxydiphenylene pyromellitamide (DOP), the model compounds of the meta and para PMDA/ODA polyamic ethyl ester, were synthesized and characterized by NMR and FTIR spectroscopy. Investigation of the imidization in d6-DMSO solution using NMR and FTIR techniques has shown that both the half imide and imide were formed. Quantitative analysis of the curing rates and degrees of conversion of the isomers in dilute d6-DMSO solution as a function of time under isothermal conditions or as function of temperature at fixed time (1 h) indicated that the kinetics of the ring closure reaction of the meta and para isomers were the same within 10%. This suggests that intrinsic reactivity differences between the isomers do not have much effect on the imidization process and do not account for the differences in rate that have been observed for the meta and para polymers in the solid state. No interconversion between the two isomeric forms occurred below 180°C, as has been observed for polyamic acids and their model compounds. The degree of conversion strongly depended on the reaction temperature and increased quickly after 170°C. The rate constant of the second ring closure reaction was found to be approximately three to four times the rate constant of the first ring closure reaction. © 1996 John Wiley & Sons, Inc. 相似文献
5.
Qiujin Zhang Weixin Lin Guang Yang Qun Chen 《Journal of Polymer Science.Polymer Physics》2002,40(19):2199-2207
The phase structure of a series of ethylene‐vinyl acetate copolymers has been investigated by solid‐state wide‐line 1H NMR and solid‐state high‐resolution 13C NMR spectroscopy. Not only the degree of crystallinity but the relative contents of the monoclinic and orthorhombic crystals within the crystalline region varied with the vinyl acetate (VA) content. Biexponential 13C NMR spin–lattice relaxation behavior was observed for the crystalline region of all samples. The component with longer 13C NMR spin–lattice relaxation time (T1) was attributed to the internal part of the crystalline region, whereas the component with shorter 13C NMR T1 to the mobile crystalline component was located between the noncrystalline region and the internal part of the crystalline region. The content of the mobile crystalline component relative to the internal part of the crystalline region increased with the VA content, showing that the 13C NMR spin–lattice relaxation behavior is closely related to the crystalline structure of the copolymers. © 2002 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 40: 2199–2207, 2002 相似文献
6.
7.
Guo-Ping Chang-Chien 《Journal of polymer science. Part A, Polymer chemistry》1998,36(16):2849-2863
A series of new and high-purity hydrocarbon liquid crystal monomers were synthesized through the acylation reaction, deoxygenation reaction, and Grignard reaction. 1H-NMR spectra and elemental analyses were used to examine their purity. The liquid crystalline polysiloxane polymers were obtained by grafting the monomers onto poly(methylhydrosiloxane). The thermal transition temperature, mesomorphic properties, and mesophase textures of the monomers and the polymers were determined by differential scanning calorimetry (dsc), polarized optical microscopy, and X-ray diffraction analysis. Moreover, we observed the even–odd effect of the smectic/isotropic transition temperature with the length variation of the substituents. In this study, we found by X-ray diffraction that the liquid crystalline polysiloxane polymers undergo a transition from smectic B to smectic E mesophase. However, dsc has difficulty detecting the phase transition process. By considering the spin–lattice relaxation time (T1), we can systematically explain the relation between the flexibility of the substituent with the smectic/isotropic transition temperature. © 1998 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 36: 2849–2863, 1998 相似文献
8.
《Magnetic resonance in chemistry : MRC》2003,41(7):489-502
Biological reactions are mostly concerned with selective interactions between small ligands and macromolecular receptors. The same ligands may activate responses of different intensities and/or effects in the presence of different receptors. Many approaches based on spectroscopic and non‐spectroscopic methods have been used to study interactions between small ligands and macromolecular receptors, including methods based on NMR and IR spectroscopic analysis of the solution behaviour of the ligand in the presence of receptors. In this work, we investigated the interaction between ovine serum albumin with two amphenicolic antibiotics [chloramphenicol (CAP) and thiamphenicol (TAP)], using a combined approach based on NMR and IR methodologies, furnishing complementary information about the recognition process occurring within the two systems. The two ligands, despite their similar structures, showed different affinities towards albumin. NMR methodology is based on the comparison of selective ( ) and non‐selective ( ) spin–lattice relaxation rates of the ligands in the presence and absence of macromolecular receptors and and temperature dependence analysis. From these studies, the ligand–receptor binding strength was evaluated on the basis of the ‘affinity index.’ The derivation of the affinity index from chemical equilibrium kinetics for both the CAP–albumin and TAP–albumin systems allowed a comparison of the abilities of the two amphenicolic antibiotics to interact with the protein. IR methodology is based on the comparison of the ligand–protein ‘complex’ spectra with those of the non‐interacting systems. On the basis of the differences revealed, a more thorough IR analysis was performed in order to understand the structural changes which occurred on both ligand and protein molecules within the interacting system. Copyright © 2003 John Wiley & Sons, Ltd. 相似文献
9.
Volker Friebolin Silvia Marten Klaus Albert 《Magnetic resonance in chemistry : MRC》2010,48(2):111-116
In the current work a racemate of (R)‐ and (S)‐benzylmandelate was separated with a stereoselective polysaccharide‐based chiral stationary phase by HPLC. To elucidate the occurring chiral molecular recognition processes in the heterogeneous system used, NMR spectroscopy was chosen under high resolution/magic angle spinning (HR/MAS) NMR conditions in the suspended state. Therefore, and as a proof of concept, a combination of several NMR methods such as spin–lattice relaxation time (T1) measurements (T1), the saturation transfer difference, and the 2D experiment of the transferred nuclear overhauser enhancement spectroscopy technique were applied. With HR/MAS NMR it is feasible to combine NMR and chromatography to achieve further insights into the separation process. Copyright © 2009 John Wiley & Sons, Ltd. 相似文献
10.
The conversion of poly(amic acid) into polyimide (PI) was achieved with far‐infrared radiation (FIR) and conventional thermal treatments. The structure and properties of PI films during different stages of imidization were studied with Fourier transform infrared spectroscopy, weight‐loss analysis during imidization, tensile property measurements, and dynamic mechanical thermal analysis. The effects of the imidization degree, postimidization, and solvent on the thermal and mechanical properties of PI films were quantitatively investigated. The corresponding structural changes were also examined. The experimental results showed that the imidization process proceeded more quickly and more completely in an FIR oven than in a conventional oven. A prolonged FIR treatment at a lower temperature (25–100 °C) accelerated the imidization process. The tensile stress–strain curves had a fanlike distribution with the development of the FIR imidization process and a fishtail distribution with conventional thermal imidization. During FIR imidization, the best tensile properties were obtained at 340 °C, and thermooxidative degradation occurred at about 420 °C. © 2004 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 42: 2490–2501, 2004 相似文献
11.
《Magnetic resonance in chemistry : MRC》2002,40(11):733-737
From the reaction of 2‐acetylcyclopentanone with benzoylhydrazine, the cyclopentanopyrazole derivative 4 was obtained as the only product. The structural assignment of this compound and also of the hydroxypyrazoline 3 , used as a model compound, was established by analysis of their NMR spectra (1H, 13C, COSY, NOESY, HETCOR and COLOC). Copyright © 2002 John Wiley & Sons, Ltd. 相似文献
12.
Mumdooh Ahmed Alexander Marchanka Teresa Carlomagno 《Angewandte Chemie (International ed. in English)》2020,59(17):6866-6873
Solid‐state NMR (ssNMR) is applicable to high molecular‐weight (MW) protein assemblies in a non‐amorphous precipitate. The technique yields atomic resolution structural information on both soluble and insoluble particles without limitations of MW or requirement of crystals. Herein, we propose and demonstrate an approach that yields the structure of protein–RNA complexes (RNP) solely from ssNMR data. Instead of using low‐sensitivity magnetization transfer steps between heteronuclei of the protein and the RNA, we measure paramagnetic relaxation enhancement effects elicited on the RNA by a paramagnetic tag coupled to the protein. We demonstrate that this data, together with chemical‐shift‐perturbation data, yields an accurate structure of an RNP complex, starting from the bound structures of its components. The possibility of characterizing protein–RNA interactions by ssNMR may enable applications to large RNP complexes, whose structures are not accessible by other methods. 相似文献
13.
Uffe Anthoni Carsten Christophersen Claus Flensburg Mogens H. Jakobsen Jan Jensen Per H. Nielsen 《Structural chemistry》1996,7(2):103-110
The solid-state structure of Boc-Gly-Trp-Ala-OBut was determined by single-crystal X-ray diffraction analysis. The tripeptide gave crystals belonging to the orthorhombic systemP212121 and at 122.0(5) K:a=11.0663(12),b=14.107(2),c=17.275(2) Å,V=2697.0(5) Å3Z=4,R(F)=0.0259, andR
w(F)=0.0695. The peptide adopts a type-I-turn in the solid state with a single, rather weak, intramolecular hydrogen bond between the Boc-CO and Ala-NH groups (NO 3.082(1) Å, <NHO 167(1)°). The conformation of the Boc-Gly-Trp-Ala-OBut peptide has also been studied by1H NMR spectroscopy. The solvent and temperature dependencies of NH chemical shifts suggests that this hydrogen bond is broken and that all amide protons are solvent exposed in CDCl3 and (CD3)2SO. 相似文献
14.
Sergey V. Fedorov Leonid B. Krivdin Yury Yu. Rusakov Igor A. Ushakov Natalia V. Istomina Natalia A. Belogorlova Svetlana F. Malysheva Nina K. Gusarova Boris A. Trofimov 《Magnetic resonance in chemistry : MRC》2009,47(4):288-299
Theoretical energy‐based conformational analysis of bis(2‐phenethyl)vinylphosphine and related phosphine oxide, sulfide and selenide synthesized from available secondary phosphine chalcogenides and vinyl sulfoxides is performed at the MP2/6‐311G** level to study stereochemical behavior of their 31P–1H spin–spin coupling constants measured experimentally and calculated at different levels of theory. All four title compounds are shown to exist in the equilibrium mixture of two conformers: major planar s‐cis and minor orthogonal ones, while 31P–1 H spin–spin coupling constants under study are found to demonstrate marked stereochemical dependences with respect to the geometry of the coupling pathways, and to the internal rotation of the vinyl group around the P(X)‐C bonds (X = LP, O, S and Se), opening a new guide in the conformational studies of unsaturated phosphines and phosphine chalcogenides. Copyright © 2009 John Wiley & Sons, Ltd. 相似文献
15.
NMR measurements ((119)Sn chemical shift, line width and (13)C relaxation) were made on triphenyltin chloride in two solutions, 2.5 and 0.75 mol% in CDCl(3), at several temperatures. The (13)C spin-lattice relaxation time and NOE data for the phenyl carbons provide the corresponding correlation times for the overall molecular reorientational motion and the internal rotation of the phenyl groups. The results are indicative of a weak intermolecular association of the triphenyltin chloride molecules in solution and are discussed with reference to a model for intermolecular phenyl ring π-π stacking interactions. 相似文献
16.
Dorota Kowalczuk Agata Gadysz Monika Pitucha Daniel M. Kamiski Agnieszka Baraska Bartomiej Drop 《Molecules (Basel, Switzerland)》2021,26(5)
Bacterial strains become resistant to almost all classes of antibiotics, which makes it necessary to look for new substitutes. The non-absorbable ciprofloxacin–biguanide bismuth complex, used locally, may be a good alternative to a conventional therapy. The purpose of this study was to study the structure of the proposed ciprofloxacin (CIP) -bismuth(III)—chlorhexidine (CHX) composite (CIP-Bi-CHX). The spectroscopic techniques such as UV-VIS (ultraviolet-visible) spectroscopy, FTIR (Fourier-transform infrared) spectroscopy and NMR (Nuclear Magnetic Resonance) spectroscopy were used for structure characterization of the hybrid compound. The performed analysis confirmed the presence of the two active components—CIP and CHX and revealed the possible coordination sites of the ligands with bismuth ion in the metallo-organic structure. Spectroscopic study showed that the complexation between Bi(III) and CIP occurs through the carboxylate and ketone groups of the quinolone ring, while CHX combines with the central ion via the biguanide moieties. 相似文献
17.
Yury Yu. Rusakov Leonid B. Krivdin Maxim V. Penzik Vladimir A. Potapov Svetlana V. Amosova 《Magnetic resonance in chemistry : MRC》2012,50(10):653-658
Stereochemical structure of nine Z‐2‐(vinylsulfanyl)ethenylselanyl organyl sulfides has been investigated by means of experimental measurements and second‐order polarization propagator approach calculations of their 1H–1H, 13C–1H, and 77Se–1H spin–spin coupling constants together with a theoretical conformational analysis performed at the MP2/6‐311G** level. All nine compounds were shown to adopt the preferable skewed s‐cis conformation of their terminal vinylsulfanyl group, whereas the favorable rotational conformations with respect to the internal rotations around the C–S and C–Se bonds of the internal ethenyl group are both skewed s‐trans. Stereochemical trends of 77Se–1H spin–spin coupling constants originating in the geometry of their coupling pathways and the selenium lone pair effect were rationalized in terms of the natural J‐coupling analysis within the framework of the natural bond orbital approach. Copyright © 2012 John Wiley & Sons, Ltd. 相似文献
18.
Calculations of 1 JNH, 1h JNH and 2h JNN spin–spin coupling constants of 27 complexes presenting N–H·N hydrogen bonds have allowed to analyze these through hydrogen‐bond coupling as a function of the hybridization of both nitrogen atoms and the charge (+1, 0, ? 1) of the complex. The main conclusions are that the hybridization of N atom of the hydrogen bond donor is much more important than that of the hydrogen bond acceptor. Positive and negative charges (cationic and anionic complexes) exert opposite effects while the effect of the transition states ‘proton‐in‐the‐middle’ is considerable. Copyright © 2008 John Wiley & Sons, Ltd. 相似文献
19.
20.
Hexamethyl-1,2,3-tristanna-[3]ferrocenophane ( 1 ) was prepared by the reaction of 1,1′-bis(dimethylstannyl)ferrocene ( 3 ) with bis(diethylamino)dimethylstannane. The molecular structure of 1 was determined by X-ray crystallography. The monoclinic unit cell (space group P21/c; a = 18.659(4), b = 17.311(3), c = 13.719(3) Å; β = 111.02(3)°) contains two independent molecules which differ slightly in their conformation. The cyclopentadienyl rings are almost parallel, but the positions of the substituted carbon atoms are twisted by τ £ 62° with respect to the ecliptic positions. The reactivity of 1 towards iodine and chalcogens E (E = S, Se, Te) was studied. Iodine reacts to give 1,1′-bis[iodo(dimethyl)stannyl]ferrocene ( 6 ) and dimethyltin diiodide. In the case of the chalcogens, the detectable and isolated products are 1,3-distanna-2-chalcogena-[3]ferrocenophanes (E = S ( 7 ), Se ( 8 ), Te ( 9 )) in addition to trimeric dimethyltin chalcogenides, (Me2SnE)3. Crystals suitable for X-ray structural analysis could be obtained of 1,3-distanna-2-thia-[3]ferrocenophane ( 7 ); the triclinic unit cell (space group P 1) has the dimensions a = 6.538(2), b = 9.013(2), c = 15.442(2) Å; α = 92.15(2), β = 91.89(2), γ = 109.43(2)°. The molecular structures of 1 and 7 are compared with those of other 1,3-distanna-[3]ferrocenophanes. All compounds were studied by NMR spectroscopy (1H, 13C, 77Se, 119Sn and 125Te NMR) in order to establish the presence of the [3]ferrocenophanes 7 – 9 and of the cycles (Me2SnE)3 in solution. 相似文献