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 15 毫秒
1.
The conformational flexibility and dynamics of two (1→6)‐linked disaccharides that are related to the action of the glycosyl transferase GnT‐V have been investigated. NMR NOE and T‐ROE spectroscopy experiments, conformation‐dependent coupling constants and molecular dynamics (MD) simulations were used in the analyses. To facilitate these studies, the compounds were synthesised as α‐d‐ [6‐13C]‐Manp‐OMe derivatives, which reduced the 1H NMR spectral overlap and facilitated the determination of two‐ and three‐bond 1H,1H, 1H,13C and 13C,13C‐coupling constants. The population distribution for the glycosidic ω torsion angle in α‐d‐ Manp‐(1→6)‐α‐d‐ Manp‐OMe for gt/gg/tg was equal to 45:50:5, whereas in α‐d‐ Manp‐OMe it was determined to be 56:36:8. The dynamic model that was generated for β‐d‐ GlcpNAc‐(1→6)‐α‐d‐ Manp‐OMe by MD simulations employing the PARM22/SU01 CHARMM‐based force field was in very good agreement with experimental observations. β‐d‐ GlcpNAc‐(1→6)‐α‐d‐ Manp‐OMe is described by an equilibrium of populated states in which the ? torsion angle has the exo‐anomeric conformation, the ψ torsion angle an extended antiperiplanar conformation and the ω torsion angle a distribution of populations predominantly between the gauchetrans and the gauchegauche conformational states (i.e., gt/gg/tg) is equal to 60:35:5, respectively. The use of site‐specific 13C labelling in these disaccharides leads to increased spectral dispersion, thereby making NMR spectroscopy based conformational analysis possible that otherwise might be difficult to attain.  相似文献   

2.
d ‐Glucaric acid (GA) is an aldaric acid and consists of an asymmetric acyclic sugar backbone with a carboxyl group positioned at either end of its structure (i.e., the C1 and C6 positions). The purpose of this study was to conduct a conformation analysis of flexible GA as a solution in deuterium oxide by NMR spectroscopy, based on J‐resolved conformation analysis using proton–proton (3JHH) and proton–carbon (2JCH and 3JCH) coupling constants, as well as nuclear overhauser effect spectroscopy (NOESY). The 2JCH and 3JCH coupling constants were measured using the J‐resolved heteronuclear multiple bond correlation (HMBC) NMR technique. NOESY correlation experiments indicated that H2 and H5 were in close proximity, despite the fact that these protons were separated by too large distance in the fully extended form of the chain structure to provide a NOESY correlation. The validities of the three possible conformers along the three different bonds (i.e., C2? C3, C3? C4, and C4? C5) were evaluated sequentially based on the J‐coupling values and the NOESY correlations. The results of these analyses suggested that there were three dominant conformers of GA, including conformer 1 , which was H2H3:gauche, H3H4:anti, and H4H5:gauche; conformer 2 , which was H2H3:gauche, H3H4:anti, and H4H5:anti; and conformer 3 , which was H2H3:gauche, H3H4: gauche, and H4H5:anti. These results also suggested that all three of these conformers exist in equilibrium with each other. Lastly, the results of the current study suggested that the conformational structures of GA in solution were ‘bent’ rather than being fully extended. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

3.
Conformational analysis of di-ortho-substituted diphenylmethanes and Biphenyl ethers containing I, CCH, and CCCCH substituents was carried out by the molecular mechanics method using the MM3 program. Several minima on the potential energy surface, which correspond to thegg, gt, tg, andort conformations, were found. An increase in the length of the linear substituent results in a substantial decrease in the difference in the relative energies of conformers. Barriers to conformational transitions between thegt, tg, andort conformers are less than 2 kcal mol–1. The transitionort-gg requires expenditure of energy of up to 5 kcal mol–1. Two valleys of centrosymmetric pairs of thegt, tg, andort conformers are separated by a barrier of up to 6 kcal mol–1.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No, 12, pp, 2882–2885, December, 1996.  相似文献   

4.
The self‐complementary UA and AU dinucleotide analogues 41 – 45, 47, 48 , and 51 – 60 were prepared by Sonogashira coupling of 6‐iodouridines with C(5′)‐ethynylated adenosines and of 8‐iodoadenosines with C(5′)‐ethynylated uridines. The dinucleotide analogues associate in CDCl3 solution. The C(6/I)‐unsubstituted AU dimers 51 and 54 prefer an anti‐oriented uracilyl group and form stretched linear duplexes. The UA propargyl alcohols 41 and 43 – 45 possess a persistent intramolecular O(5′/I)? H???N(3/I) H‐bond and, thus, a syn‐oriented adeninyl and a gt‐ or tg‐oriented ethynyl moiety; they form corrugated linear duplexes. All other dimers form cyclic duplexes characterized by syn‐oriented nucleobases. The preferred orientation of the ethynyl moiety (the C(4′),C(5′) torsion angle) defines a conformation between gg and one where the ethynyl group eclipses O(4′/I). The UA dimers 42, 47 , and 48 form Watson–Crick H‐bonds, the AU dimers 56 and 58 – 60 H‐bonds of the Watson–Crick‐type, the AU dimers 53 and 55 reverse‐Hoogsteen, and 57 Hoogsteen H‐bonds. The pairing mode depends on the substituent of C(5′/I) (H, OSiiPr3; OH) and on the H‐bonds of HO? C(5′/I) in the AU dimers. Association constants were derived from the concentration‐dependent chemical shift for HN(3) of the uracilyl moiety; they vary from 45–104 M ?1 for linear duplexes to 197–2307 M ?1 for cyclic duplexes. The thermodynamic parameters were determined by van't Hoff analysis of the temperature‐dependence of the (concentration‐dependent) chemical shift for HN(3) of the uracilyl moiety. Neglecting stacking energies, one finds an average energy of 3.5–4.0 kcal/mol per intermolecular H‐bond. Base stacking is evidenced by the temperature‐dependent CD spectra. The crystal structure of 54 shows two antiparallel chains of dimers connected by Watson‐Crick H‐bonds. The chains are bridged by a strong H‐bond between the propargylic OH and O?C(4) and by weak reverse A ? A Hoogsteen H‐bonds.  相似文献   

5.
Some monomer model compounds of lignin have been selectively 2H and 13C labelled: vanillin, ethyl ferulate, coniferyl alcohol and ethyl hydrogen malonate. Deuterium isotope effects on the 13C chemical shifts in [formyl-2H]vanillin, [5-2H]vanillin and [α,α,5-2H3]coniferyl alcohol made the unambiguous assignment of the aromatic 13C signals possible. Absolute 1,2,3J(CC) values have been determined on 13C spectra of [formyl-13C]vanillin, and of ethyl ferulate and coniferyl alcohol in which the vinylic C-γ and C-β carbons were 13C enriched. It has been possible to measure 4J(C?O, C-4) in vanillin and 4J(C-γ, C-4) in ethyl ferulate. The determination of 1,2,3,4J (CH) absolute values was done by means of gated decoupled 13C spectra of the non-labelled compounds. When second order effects made the use of this technique impossible we determined certain J(CH) values and their signs either by analysing the 1H NMR spectra of 13C labelled coniferyl alcohol [2J(C-β, H-γ), 2J(C-β, H-α), 2J(C-γ, H-β), 3J(C-γ, H-α)] or by a double irradiation experiment on the 250 MHz 1H NMR spectrum of ethyl [β-13C] ferulate [for 2J(C-β, H-γ)].  相似文献   

6.
The conformational behavior of cellobiose (D -glc-ß(1→4)-D -glc), cellotetraose, and cellooctaose was studied by a combination of energy minimization and molecular dynamics simulations in vacuo at 400 K. These diand oligosaccharide models have considerable flexibility and exhibit a variety of different motions in glycosidic and exocyclic torsions. The glycosidic ?, ψ torsions moved frequently between two local minima on the cellobiose energy surface in the region of known crystal structures. Transitions of the hydroxymethyl side chain were observed between gt,gg, and tg conformations accompanied by changes in intramolecular hydrogen bonding patterns. A reasonable fit to the experimental optical rotation and nuclear magnetic resonance vicinal coupling data of cellobiose in solution required a distribution of its conformations. The oligomers, although generally extended, assumed a more coiled or twisted shape than is observed in the crystalline state of cellulose and exhibited considerable backbone motion due to local ring rotations about the glycosidic bonds. Long-lived transitions to structures having torsion angles 180° from the major minima (ring flips) introduced kinks and bends into the tetramer and octamer. While the glucose rings of the structures remained primarily in the 4C1 conformation, twist and boat structures were also observed in the tetramer and octamer structures. Reducing the simulation temperature to 300 K eliminated some of the transitions seen at 400 K. © 1993 John Wiley & Sons, Inc.  相似文献   

7.
Summary Addition of sulfur dichloride to tetrachlorocatechol-bisallylether (1) yields the 9- and 10-ring thia crown ether derivatives2 and3, respectively, together with the dithia-18-crown-6-ether4. The 10-membered ring compound3 represents the first thia macrocycle containing bothMarkovnikov andanti-Markovnikov constitution of the -chloro-thio structural segments in the same molecule. By1H and13C NMR spectroscopy, equal amounts of two preferred conformers of the only isolated diastereomer of3 were observed at temperatures below –50°C. The signals were assigned to these conformers using COSY, HETCOR, and phase sensitive NOESY spectra at low temperatures. The preferred conformations and the relative configuration were determined using the different effects of gauche -and anti -positions in13C NMR chemical shifts and analyzing vicinal3 J H,H coupling constants. These results were confirmed by molecular mechanics calculations.Dedicated to Prof. Dr.Rolf Borsdorf on the occasion of his 65th birthday  相似文献   

8.
Abstract

The rotational freedom of the C5/C6 bond of hexopyranosides very often governs the conformation of oligosaccharides, especially when 1→6 linkages are present in the saccharide. Conformational analysis may be complicated by the existence of a dynamic equilibrium between three staggered rotamers in the C5/C6 fragment.1 The population of the individual rotamers can be deduced from J5,6s and J5,6R coupling constants by ap- plying an equation that connects the experimentally derived time-averaged coupling constants to the populational equilibrium. A prerequisite for this approach is the assign- ment of the prochiral protons H-6R and H-6S2 and a correct evaluation of the effects of substituents on the coupling constant.  相似文献   

9.
Abstract

1,2-O-Isopropylidene-α-D-xylo-hexofuranos-5-ulose (2) was deprotected in aqueous acid solution to give a mixture of at least six isomeric forms and one anhydro form of the parent ketoaldohexose, D-xylo-hexos-5-ulose (3), commonly referred to as 5-keto-glucose. Structural assignment of each form was made based on high field 1H and 13C NMR studies of the mixture in aqueous (D2O) solution. The dominant isomeric form of 3 was observed to have the pyranose structure 1R,5R-D-xlyo-hexo-pyranos-5-ulose (3a, 67 %) with the next most abundant form being an anhydro structure, 1S,5S-l,6-anhydro-D-xylo-hexopyranos-5-ulose (3c, 18 %). Included among the other isomers were the a and β-1,4-furanose (3d, 3e) and 1-aldehydrol β-5,2-furanose (3f) structures. The isomer present in least amount (3g, > 1 %) is assigned as the α-anomer of 3f. Experimentally determined JC-1,H-1 values were useful in support of assigned isomer structures.  相似文献   

10.
Abstract

The anomeric 1JC,J value has for some time been successfully used for structural elucidation of oligosaccharides. The first recorded spectra on unenriched carbohydrates were published by Bock et al. in 1973 and pointed out that the anomeric 1JC,J value could be used for assignment of the anomeric configuration since pyranoses with an axial H-1 have a 1JC,J value which is approximately 10 Hz lower than the corresponding value in compounds with an equatorial H-1.1–3 Their method recording non decoupled 13C NMR spectra had the disadvantage of being insensitive and thereby time-consuming. Introduction of INEPT4 followed by technical developments, has overcome this disadvantage. Nowadays, with reverse detection techniques, technical skill instead of sensitivity is a limiting factor. In this communication I would like to emphasize that the anomeric 1JC,J value can be used for detecting 1,2-orthoester formation as well as for establishing a- or P-configuration.  相似文献   

11.
The solution conformation of two cyclic-casomorphin-5 analogues H-Tyr-c(-d-Orn-Phe-Pro-Gly-)1 and H-Tyr-c(-Orn-Phe-Pro-Gly-)2 in DMSO-d 6 was studied by NMR spectroscopy and accompanying force field calculations. By especially employing1H,13C, and15N chemical shifts, respectively, the temperature coefficient of the amide proton chemical shifts,3 J NH,CH andJ CH.CH coupling constants, respectively, and nuclear Overhauser effects in the rotating frame (ROEs), in the case of1, only one preferred conformer could be identified. In the case of2, two or even more preferred conformers were found, readily interconverting on the NMR time scale. Empirical force field calculations using the SYBYL 6.0 software (TRIPOS) corroborate the experimental NMR results obtained. The conformational behavior of the compounds studied is discussed with respect to the receptor specificity of the-casomorphins studied.  相似文献   

12.
A mixture of stereoisomers of 2,4-dimethoxybicyclo[3.3.1]nonan-9-one was prepared, separated by column chromatography and characterized by 60 MHz 1H NMR spectroscopy using Eu(fod)3. A double chair conformation with axial methoxyl groups is established for (1R,2S,4R,5S)-2,4-dimethoxybicyclo[3.3.1]-nonan-9-one on the basis of the J(12), J(2,H-3 exo) and J(2,3 endo) values and the chemical shifts for H-2(4). The conformation of some related compounds is subsequently inferred.  相似文献   

13.
Proton–proton 3J, 4J and 5J NMR coupling constants have been calculated for cyclohexane and monosubstituted cyclohexane conformers (substitiuents: Li, CH3, OH, F) by the two methods mentioned. Comparing the two methods on the basis of group theory, we show the necessity to use the second. The results from this method are compared with those of the literature.  相似文献   

14.
Synthesis of 2-Substituted Imidazole Nucleosides Condensation of the trimethylsilyl derivatives of 2-substituted diethyl and dimethyl imidazole-4,5-dicarboxylates ( 3–5 and 7–9 ) with 1-O-acetyl-2,3,5-tri-O-benzoyl-β-D -ribofuranose ( 2 ) in the presence of trimethysilyl trifluoromethanesulfonate provided the 2-substituted diethyl and dimethyl 1-(2′,3′, 5′-tri-O-benzoyl-β-D -ribofuranosyl)imidazole-4, 5-dicarboxylates 10–15 . These were treated with ammonia to afford the 2-substituted 1-(β-D -ribofuranosyl)imidazole-4,5-dicarboxamides 16–21 . Treatment of 2-methyl-( 16 ) and 2-ethyl-1-(β-D -ribofuranosyl)imidazole-4,5-dicarboxamide ( 17 ) with fuming nitric acid in oleum at ?30° yielded the nitric acid esters 23 and 24 . Besides the esterification of the sugar hydroxyl groups one H-atom of the imidazolecarboxamide function at C(5) in these nucleosides was also substituted by the NO2 group. The conformations in solution of 16 and 23 have been determined by 1H- and 13C-NMR. spectroscopy. These studies indicate that the nucleosides exist in dimethyl-sulfoxide solution preferentially in the S-gg-syn-conformation ( 16 ) and N-gt-conformation ( 23 ). In the crystal structure of nucleoside 23 , the ribose was found to be in the O(1′)endo, C(1′)exo twist conformation. The conformation about C(4′), C(5′) is gauche-trans and the molecule exists in the syn form.  相似文献   

15.
Bis(5,5-dimethyl-3-hydrazonocyclohex-1-enyl) sulfide was synthesized by the reaction of bis(5,5-dimethyl-3-thioxocyclohex-1-enyl) sulfide or its oxygen analog with hydrazine. Conformational lability of the molecule of dihydrazone was studied by the DFT(B3LYP) method using expanded basis sets 6–311G and 6–11G(d,p). Analysis of vibration IR and Raman spectra of the most stable conformers of the isolated molecule of dihydrazone was performed at the harmonic approximation. Using the Onsager SCRF solvation model the absence of solvent effect on the relative stability of the conformers was shown. The photoconductivity of dihydrazone was studied. Low value of the ratio of photocurrent to dark current (J p/J d = 2.5–3.5) for dyhydrazone was assigned to the lability of its stereoelectronic structure, which was in line with the data of 1H, 13C, and 15N NMR spectra.  相似文献   

16.
Periodic and molecular cluster density functional theory calculations were performed on the Iα (001), Iα (021), Iβ (100), and Iβ (110) surfaces of cellulose with and without explicit H2O molecules of hydration. The energy-minimized H-bonding structures, water adsorption energies, vibrational spectra, and 13C NMR chemical shifts are discussed. The H-bonded structures and water adsorption energies (ΔEads) are used to distinguish hydrophobic and hydrophilic cellulose–water interactions. O–H stretching vibrational modes are assigned for hydrated and dry cellulose surfaces. Calculations of the 13C NMR chemical shifts for the C4 and C6 surface atoms demonstrate that these δ13C4 and δ13C6 values can be upfield shifted from the bulk values as observed without rotation of the hydroxymethyl groups from the bulk tg conformation to the gt conformation as previously assumed.  相似文献   

17.
The conformational behavior of the capsular polysaccharide obtained from a fast‐growing soybean‐nodulating rhizobia (strain B33) isolated from Xinjiang Autonomous Region (Eastern China) has been analyzed by NMR and molecular mechanics simulations. This polysaccharide has the repeating unit →6)‐4‐O‐Me‐α‐d‐Glcp‐(1→4)‐3‐O‐Me‐β‐d‐GlcpA‐(1→. The NMR results indicate that the α‐(1→4) linkage may adopt a variety of conformations, and that at least two of the resulting minima must exist in solution. NOE data agree with an 85:10:5 ratio for the lowest‐energy conformations. In the case of the β‐(1→6) linkage, NMR indicates that the rotamer gg is highly populated. Experimental and calculated NOE intensities match well when the global energy minimum conformation for this linkage has exclusively the gg orientation. The influence of the adjacent methyl group on the glycosyloxymethyl population has been evaluated by simulation of a disaccharide without this group. A relative destabilization of gt rotamer has been found. Long chains have been simulated using a Metropolis algorithm at different ratios of the gg and gt rotamers in the glucose moiety. It was observed that the increase in population of the gt rotamer yielded more close contacts in the chain.  相似文献   

18.
All J(P? H) and J(P? C) values, including signs, have been obtained in acetylenic and propynylic phosphorus derivatives, R2P(X)? C?C? H and R2P(X)? C?C? CH3 (X ? oxygen, lone pair and R ? C6H5, N(CH3)2, OC2H5, N(C6H5)2, Cl) from 1H and 13C NMR spectra. In PIV derivatives the following signs are obtained: 1J(P? C)+, 2J(P? C)+, 3J(P? C)+, 3J(P? H)+, 4J(P? H)? . Linear relations are observed between 1J(P? C), 2J(P? C) and 3J(P? C) versus 3J(P? H), indicating that these coupling constants are mainly dependent on the Fermi contact term, though the other terms of the Ramsey theory do not seem to be negligible for 1J(P? C) and 2J(P? C). In PIII derivatives these signs are: 1J(P? C)- and +, 2J(P? C)+, 3J(P? C)-, 3J(P? H)-, 4J(P? H)+. Only 3J(P? C) and 3J(P? H) reflect a small contribution of the Fermi contact term while in 1J(P? C) and 2J(P? C) this contribution seems to be negligible relative to the orbital and/or spin dipolar coupling mechanisms.  相似文献   

19.
Quantum chemical study of conformational isomerization of 2-methyl-5-alkyl- and 5-aryl-1,3,2-dioxaborinanes using RHF/6-31G(d) method led to the conclusion that the equilibrium between equatorial and axial sofa conformers is shifted to the latter form. Based on the experimental and theoretically calculated vicinal coupling constants J HH the quantitative conformational composition and the values of ΔG 0 for substituents at the C5 ring atom were established.  相似文献   

20.
Abstract

Data are presented on the magnitudes of 5 J pp and 6 J PP spin-spin coupling constants in the 31P NMR spectra of a variety of novel polyphosphite triesters.  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