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 30 毫秒
1.
Temperature coefficients (Deltadelta/DeltaT) of amide chemical shifts of N-acetylglucosamine residues have been measured in a range of oligosaccharides of the important vertebrate polysaccharide hyaluronan. Odd- and even-numbered oligosaccharides with glucuronic acid, Delta-4,5-unsaturated glucuronic acid and N-acetylglucosamine at the termini were investigated. All amide proton temperature coefficients were only slightly less negative (-6.9 to - 9.1 ppb/ degrees C) than those of amide protons in free exchange with water (approximately equal to -11 ppb/ degrees C), indicating an absence of persistent intramolecular hydrogen bonds. With the exception of amide groups in reducing-terminal N-acetylglucosamine rings, all amide proton environments have the same temperature coefficient (-6.9 ppb/ degrees C), irrespective of differences in amide group chemical shifts and (3)J(HH) coupling constants, i.e. they do not sense subtle differences in orientation of the amide group. Amide nitrogen temperature coefficients report the same phenomena but with greater sensitivity. These data provide a set of reference values for temperature coefficients measured in other carbohydrates with acetamido sugars.  相似文献   

2.
刘雪辉  林东海 《中国化学》2007,25(3):411-416
This paper describes an amide-exchange-rate-edited (AERE) NMR method that can effectively alleviate the problem of resonance overlap for proteins and peptides. This method exploits the diversity of amide proton exchange rates and consists of two complementary experiments: (1) SEA (solvent exposed amide)-type NMR experiments to map exchangeable surface residues whose amides are not involved in hydrogen bonding, and (2) presat-type NMR experiments to map solvent inaccessibly buried residues or nonexchangeable residues located in hydrogen-bonded secondary structures with properly controlled saturation transfer via amide proton exchanges with the solvent. This method separates overlapping resonances in a spectrum into two complementary spectra. The AERE-NMR method was demonstrated with a sample of ^15N/^13C/^2H(70%) labeled ribosome-inactivating protein trichosanthin of 247 residues.  相似文献   

3.
NMR spectroscopy was used to probe the conformational behavior of diastereomeric s-triazine derivatives containing two chiral amino amide substituents, in order to shed light onto the mechanism of chromatographic diastereoselectivity. Utilizing the amino hydrogen signals in the proton NMR spectrum, the population of the conformations caused by rotation about the bond between the amino nitrogen and aromatic carbon atoms could be observed. The population distribution between the three possible conformations was similar but not identical between the two diastereomers, with similar trends being observed for both bis alanine amide and bis valine amide derivatives. Based on a simple model in which it is assumed that adsorption to the hydrophobic RP-LC stationary phase occurs only for the conformations having both amino amide R-groups on the same side of the triazine ring plane, the different conformation populations between the two diastereomers obtained by NMR was consistent with the observed RP-LC elution order (L-L diastereomer followed by L-D). The predicted diastereoselectivity values from NMR data were compared to RP-LC diastereoselectivity values obtained using both C18 and polymeric columns, with both acetonitrile/water and DMSO/water mobile phases. Values obtained with the polymeric column were in better agreement with calculated values than those obtained with the C18 column, suggesting that the simple adsorption model used to calculate the diastereoselectivity is more relevant towards a simple hydrophobic polymeric surface rather than a more complex C18 stationary phase. This study indicates that proton NMR is a useful tool for studying the diastereoselective mechanism of these derivatives, due to the relatively slow C-N bond rotation caused by the significant sp(2) character of the amino nitrogen atoms.  相似文献   

4.
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   

5.
A new magic-angle spinning NMR method for measuring internuclear distances between a 13C-labeled site and amide protons is described. The magnetization of the protons evolves under homonuclear decoupling and the recoupled 13C-1H dipolar interaction, which provides simple spin-pair REDOR curves if only one 13C-labeled site is present. The modulation of the amide proton HN is detected via short 1H-15N cross polarization followed by 15N detection. The method is demonstrated on two specifically 13C- and 15N-labeled peptides, with 13C-HN distances from 2.2 to ca. 6 A. This technique promises to be particularly useful for measuring distances between 13C=O and H-15N groups, to identify hydrogen bonds in peptides and proteins.  相似文献   

6.
According to the (1)H, (13)C and (15)N NMR spectroscopic data and ab initio calculations, the strong N--H...O intramolecular hydrogen bond in the Z-isomers of 2-(2-acylethenyl)pyrroles causes the decrease in the absolute size of the (1)J(N,H) coupling constant by 2 Hz in CDCl(3) and by 4.5 Hz in DMSO-d(6), the deshielding of the proton and nitrogen by 5-6 and 15 ppm, respectively, and the lengthening of the N--H link by 0.025 A. The N--H...N intramolecular hydrogen bond in the 2(2'-pyridyl)pyrrole leads to the increase of the (1)J(N,H) coupling constant by 3 Hz, the deshielding of the proton by 1.5 ppm and the lengthening of the N--H link by 0.004 A. The C--H...N intramolecular hydrogen bond in the 1-vinyl-2-(2'-pyridyl)-pyrrole results in the increase of the (1)J(C,H) coupling constant by 5 Hz, the deshielding of the proton by 1 ppm and the shortening of the C--H link by 0.003 A. Different behavior of the coupling constants and length of the covalent links under the hydrogen bond influence originate from the different nature of the hydrogen bonding (predominantly covalent or electrostatic), which depends in turn on the geometry of the hydrogen bridge. The Fermi-contact mechanism only is responsible for the increase of the coupling constant in the case of the predominantly electrostatic hydrogen bonding, whereas both Fermi-contact and paramagnetic spin-orbital mechanisms bring about the decrease of coupling constant in the case of the predominantly covalent hydrogen bonding.  相似文献   

7.
The molecular and crystal structure of single-crystalline mesitylenesulfonic acid dihydrate (1) was determined by X-ray diffraction and IR spectroscopy. According to X-ray diffraction data, water molecules in the crystal structure form H5O2 + cations stabilized by an intracationic hydrogen bond with a length of 2.45(1) Å. The formation of the asymmetric H5O2 + cation was confirmed by IR spectroscopy. The crystallographic nonequivalence of the water molecules results in a shift of the bridging proton from the midpoint of the strong hydrogen bond in the cation toward one of the water molecules. The proton conductivity of compound 1 was measured by impedance spectroscopy. Dihydrate 1 is completely dehydrated upon prolonged storage in a dry argon glove box and undergoes the transition to the dielectric state. Compound 1 is stable in the humidity range of 32–66 rel.%. The conductivity of dihydrate 1 is (2.4±0.3) · 10?5 Ohm?1 cm?1 at 298 K, E a = 0.21±0.01 eV.  相似文献   

8.
The structures of the 1:1 hydrated proton‐transfer compounds of isonipecotamide (piperidine‐4‐carboxamide) with oxalic acid, 4‐carbamoylpiperidinium hydrogen oxalate dihydrate, C6H13N2O+·C2HO4·2H2O, (I), and with adipic acid, bis(4‐carbamoylpiperidinium) adipate dihydrate, 2C6H13N2O+·C6H8O42−·2H2O, (II), are three‐dimensional hydrogen‐bonded constructs involving several different types of enlarged water‐bridged cyclic associations. In the structure of (I), the oxalate monoanions give head‐to‐tail carboxylic acid O—H...Ocarboxyl hydrogen‐bonding interactions, forming C(5) chain substructures which extend along a. The isonipecotamide cations also give parallel chain substructures through amide N—H...O hydrogen bonds, the chains being linked across b and down c by alternating water bridges involving both carboxyl and amide O‐atom acceptors and amide and piperidinium N—H...Ocarboxyl hydrogen bonds, generating cyclic R43(10) and R32(11) motifs. In the structure of (II), the asymmetric unit comprises a piperidinium cation, half an adipate dianion, which lies across a crystallographic inversion centre, and a solvent water molecule. In the crystal structure, the two inversion‐related cations are interlinked through the two water molecules, which act as acceptors in dual amide N—H...Owater hydrogen bonds, to give a cyclic R42(8) association which is conjoined with an R44(12) motif. Further N—H...Owater, water O—H...Oamide and piperidinium N—H...Ocarboxyl hydrogen bonds give the overall three‐dimensional structure. The structures reported here further demonstrate the utility of the isonipecotamide cation as a synthon for the generation of stable hydrogen‐bonded structures. The presence of solvent water molecules in these structures is largely responsible for the non‐occurrence of the common hydrogen‐bonded amide–amide dimer, promoting instead various expanded cyclic hydrogen‐bonding motifs.  相似文献   

9.
Side-chain proton and carbon-13 resonance assignments of [13C;15N]-enriched proteins usually rely on combinations of several multi-dimensional experiments. Here, we describe a four-dimensional pulse sequence, H(C)C-COSY-TOCSY-(CACO)NH, which provides the information required to assign completely aliphatic side-chain resonance frequencies. As in widely used HCC(CO)NH-TOCSY experiments, problems due to spectral crowding are alleviated by exploiting the dispersion of backbone amide 1H and 15N signals. The modification introduced here allows signals from different side-chains to be distinguished even in the case of overlap in the 1H(N)-15N plane of the spectra. For illustration, the new method is applied to two proteins with molecular masses of 11 and 23 kDa.  相似文献   

10.
alpha, alpha-Trehalose (trehalose) is a nonreducing disaccharide of glucose and is accumulated at high concentrations in some anhydrobiotic organisms, which can survive without water for long periods and rapidly resume active metabolism upon hydration. Although it has been proposed that the intriguing mechanism of bioprotection in anhydrobiosis is conferred by a water channel, details of such a channel have yet to be revealed. We determined the crystal structure of a trehalose anhydrate to further understand the relationship between the structure of water channels and the trehalose polymorph. The space group was identical to that of the dihydrate and the lattice constants were also very similar. Among the five intermolecular hydrogen bonds between the trehalose molecules, four were preserved in the anhydrate. If dehydration of the dihydrate is slow and/or gentle enough to preserve the hydrogen bonds, transformation from the dihydrate to the anhydrate may occur. There are two different holes, hole-1 and hole-2, along one crystal axis. Hole-1 is constructed by trehalose molecules with a screw diad at its center, while hole-2 has a smaller diameter and is without a symmetry operator. Because of the screw axis at the center of hole-1, hollows are present at the side of the hole with diameters roughly equal to that of hole-1. Hole-1 and side pockets followed by hollows correspond to the positions of two water molecules of the dihydrate. The side hollows of the water channel are also observed in the water-filled hole of the dihydrate. Consequently, hole-1 is considered to be a one-dimensional water channel with side pockets. We also calculated molecular and crystal energies to examine the rapid water uptake of the anhydrate. It was demonstrated that the intermolecular interactions in the anhydrate were weaker than in the other anhydrous form, and probably also than those in amorphous trehalose. The anhydrate provides water capture for another solid form and gives protection from water uptake. These structural properties of the anhydrate may elucidate bioprotection in anhydrobiosis.  相似文献   

11.
[D ‐Leu1]‐microcystin‐LR is a recently discovered microcystin. We report the isolation of this microcystin analogue from a Microcystis aeruginosa strain isolated from the Lagoa de Iquipari, Rio de Janeiro, Brazil. The 1H and 13C NMR spectra were completely assigned in both MeOH‐d4 and DMSO‐d6. Further, the solution structure of this compound was investigated with the use of two‐dimensional NMR and the amide proton temperature dependence, and was compared with those of its analogs, microcystin‐RR and microcystin‐LR. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

12.
The N(Py).HN(amide) hydrogen bonding within the macrocyclic cavities in 9, 10, and 13 invokes their symmetrical electron-deficient structures ((1)H NMR) and consequently bind with water. This results in their poor ionophore characters. The steric requirement of methyl/benzyl substituents on amide N in 11 and 12 takes the substituents out of the cavity and thus positions the amide O toward the cavity ((1)H, (13)C NMR and X-ray analysis). This arrangement of two pyridine N and two amide O ((13)C NMR, IR) binding sites provides an appropriate environment for selective binding toward Ag(+) over Pb(2+), Tl(+), alkali, and alkaline earth cations. The increased spacer length in 14 leads to a lop-sided twist of pyridine rings (X-ray) and disturbs the above arrangement and leads to its poor binding character.  相似文献   

13.
1,6-六亚甲基二异氰酸酯自聚产物的结构表征   总被引:2,自引:0,他引:2  
用IR与NMR表征了用醋酸钾为催化剂时 1,6 六亚甲基二异氰酸酯 (HDI)自聚产物的结构 .结果表明 ,自聚主产物是三聚体异氰脲酸酯 ,主要含有三聚体异氰脲基、异氰酸根 ,同时含有由杂质带来的微量氨基甲酸酯、脲基甲酸酯基、取代脲基、缩二脲基 .一维核磁谱及二维化学位移相关谱分辨出 7种羰基 ,一种NCO基 ,确定了氮上 8种不同取代结构的分子链连接情况 .通过建立理论模型 ,定量地描述了自聚产物的结构 .  相似文献   

14.
Unambiguous and complete assignments of 1H and 13C NMR chemical shifts for five clerodane diterpenes, four of them isolated from Salvia splendens (salviarin, splendidin and splenolides A and B) and one obtained by acetylation of splenolide A, are presented. The assignments are based on 2D shift-correlated [1H,1H-COSY, 1H,13C-gHSQC-1J(C,H) and 1H,13C-gHMBC-nJ(C,H) (n=2 and 3)] and nuclear Overhauser effect (NOE) experiments. The conformation of the rings of these compounds is supported by the 3J(H,H) values and NOE results.  相似文献   

15.
Azo-hydrazone tautomeric behavior of polyazo Solophenyl red 3BL (C.I. Direct 80) dye in different solvents (water, methanol and DMSO) was investigated using 1H, 13C, NH, HH, CH COSY, HH NOESY NMR techniques and UV-vis spectroscopy. Two-dimensional NMR experiments were used to assign 1H, 13C and 15N NMR lines unambiguously. Results showed that the hydrazone-form proton NMR signal appeared in the weakest field with respect to tetramethylsilane, in comparison with the amide and phenolic proton NMR signals. UV-vis absorption spectroscopic evidences showed that azo-hydrazone mixture exists in water and DMSO solvents, but in methanol, only azo tautomer was dominant, which was in a good agreement with NMR spectroscopic results.  相似文献   

16.
The results of NMR spectroscopic and X-ray crystallographic studies are critically discussed with respect to the structure of hydantoins, their tautomerism, and their acidity. The imide NH proton of the preferred, nearly planar 2,4-imidazolidine-dione tautomer proved to be more acidic than the corresponding amide NH proton. Phenyl substituents at the ring nitrogen atoms and at C-5 are twisted from the plane of the hydantoin ring; in case ofortho substituents restricted rotation about the N-aryl bond was found and the barrier to rotation determined by dynamic NMR spectroscopy. For 5-benzyl substituents, afolded conformation of the two rings, due to intramolecular interactions, was found and for 5-exo-methylene substituted hydantoins the relevant E/Z isomerism at theexo-cyclic C, C double bond was studied. In addition, the1H and13C chemical shifts of the hydantoins proved to excellently indicate the electronic distribution along the hydantoin ring moiety. Finally, the mass spectrometric fragmentation of the hydantoins is critically discussed.Dedicated to Prof. Rolf Borsdorf on the occasion of his 65th birthday.  相似文献   

17.
Amide-amide hydrogen bonds have been implicated in directing protein folding and enhancing protein stability. Inversion transfer (13)C NMR spectroscopy and IR spectroscopy were used to compare the ability of various amide solvents and of water to alter the rate of the cis-trans isomerization of the prolyl peptide bond of Ac-Gly-[β,δ-(13)C]Pro-OMe and the amide I vibrational mode of [(13)C=O]Ac-Pro-OMe. The results indicate that secondary amides are significantly weaker hydrogen bond donors than is formamide or water. These results are most consistent with models for protein folding in which the formation of secondary structure is a cooperative process that follows hydrophobic collapse. These results also suggest that a hydrogen bond between a main-chain oxygen and an asparagine or glutamine sidechain may contribute more to protein stability than does a main-chain-main-chain hydrogen bond.  相似文献   

18.
Valinomycin is a naturally occurring cyclic dodecadepsipeptide with the formula cyclo‐[d ‐HiVA→l ‐Val →l ‐LA→l ‐Val]3 (d ‐HiVA is d ‐α‐hydroxyisovaleic acid, Val is valine and LA is lactic acid), which binds a K+ ion with high selectively. In the past, several cation‐binding modes have been revealed by X‐ray crystallography. In the K+, Rb+ and Cs+ complexes, the ester O atoms coordinate the cation with a trigonal antiprismatic geometry, while the six amide groups form intramolecular hydrogen bonds and the network that is formed has a bracelet‐like conformation (Type 1 binding). Type 2 binding is seen with the Na+ cation, in which the valinomycin molecule retains the bracelet conformation but the cations are coordinated by only three ester carbonyl groups and are not centrally located. In addition, a picrate counter‐ion and a water molecule is found at the center of the valinomycin bracelet. Type 3 binding is observed with divalent Ba2+, in which two cations are incorporated, bridged by two anions, and coordinated by amide carbonyl groups, and there are no intramolecular amide hydrogen bonds. In this paper, we present a new Type 4 cation‐binding mode, observed in valinomycin hexaaquamagnesium bis(trifluoromethanesulfonate) trihydrate, C54H90N6O18·[Mg(H2O)6](CF3SO3)2·3H2O, in which the valinomycin molecule incorporates a whole hexaaquamagnesium ion, [Mg(H2O)6]2+, via hydrogen bonding between the amide carbonyl groups and the hydrate water H atoms. In this complex, valinomycin retains the threefold symmetry observed in Type 1 binding, but the amide hydrogen‐bond network is lost; the hexaaquamagnesium cation is hydrogen bonded by six amide carbonyl groups. 1H NMR titration data is consistent with the 1:1 binding stoichiometry in acetonitrile solution. This new cation‐binding mode of binding a whole hexaaquamagnesium ion by a cyclic polypeptide is likely to have important implications for the study of metal binding with biological models under physiological conditions.  相似文献   

19.
Interaction of H3O+ or H5O2+ with 1,3-alternate tetrapropoxycalix[4]arene (1) was studied in nitrobenzene and dichloromethane using 1H and 13C NMR including transverse and rotating-frame relaxations and density functional level of theory (DFT) quantum calculations. According to NMR, the ion forms an equimolecular complex with 1 with the equilibrium constant K being 3.97 x 10(3) L.mol(-1) at 296 K. The ions are bound by strong hydrogen bonds to the phenoxy-oxygen atoms of one half of 1 and by a medium-strong hydrogen bond to the pi system of the aromatic rings of the other half. The complex appears to have C(4h) symmetry in NMR even when cooling its solution down to 213 K, which could be due either to a genuine symmetry of the complex (if the ion is H5O2+) or to fast structure averaging by ion exchange processes (if the ion is H3O+). Therefore, the dynamics of the system was studied. Using two independent NMR methods (transverse and rotating-frame relaxation), two different exchange processes were discerned with correlation times 25 x 10(-6) and 5 x 10(-6) s, the first being clearly intermolecular and the other being apparently intramolecular. The energetic aspects of the possible exchange processes were examined by DFT quantum calculations. Rotation of H3O+ ion within one binding site with the energy barrier 8.13 kcal/mol is easily possible. Intermolecular exchange by freeing the ion from the complex has too high a barrier but cooperative interaction of the ion with additional water molecules makes it viable. The intramolecular exchange (or hopping) of the H3O+ ion between the two sites of the molecule is not viable in the classical manner, the barrier being 25.6 kcal/mol. Quantum tunneling of the ion is highly improbable, too. Alternative mechanisms including concerted two-ion intermolecular exchange and cooperative interaction with another bound water molecule including complexation with proton dihydrate H5O2+ are discussed.  相似文献   

20.
本文测定了10个乙酰苯胺衍生物的~1H,~(13)C NMR谱,采用质子自旋去偶,质子选择去偶,质子偏共振去偶,质子去偶的~(13)C DEPT和选择INEPT等技术对其谱线进行归属。系统地研究了围绕羰基碳氮键的受阻旋转,确认了该类化合物在溶液中存在E和Z型构象体,并探讨了各种取代基对形成异构体比例的影响。  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