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.
Ab initio MO calculations were carried out at the MP2/6-311++G(d,p) level to investigate the conformational energy of 2-substituted oxanes and 1,3-dioxanes. It has been found that the Gibbs free energies of the axial conformers are smaller than those of the corresponding equatorial conformers in every case when the 2-substituent Z is electron withdrawing (OCH(3), F, Cl, Br). The difference in Gibbs energy between the equatorial and axial conformers DeltaG(eq-ax) increases from Z=OCH(3) to F, Cl, and then to Br. In the axial conformers, the interatomic distance between Z and the axial C-H, separated by four covalent bonds, has been found to be appreciably shorter than the van der Waals distance, suggesting the importance of the five-membered CH/n (CH/O or CH/halogen) hydrogen bond in stabilizing these conformations. Natural bonding orbital (NBO) charges of the relevant atoms have been shown to be different between the two conformers: more positive for H and more negative for C in the axial conformers than in the corresponding equatorial conformers. In view of the above findings, we suggest that the CH/n hydrogen bond plays an important role in stabilizing the axial conformation in 2-substituted oxanes and 1,3-dioxanes, and by implication, in the anomeric effect in carbohydrate chemistry.  相似文献   

2.
2-Methoxytetrahydropyran (1), -thiopyran (2) and -selenopyran (3) have been chosen as model compounds to investigate the origin of the anomeric effect (AE). The impacts of the hyperconjugation, electrostatic and steric interactions on the conformational preferences of compounds 13 have been analysed by means of complete basis set-4, hybrid-density functional theory (B3LYP/6-311+G**) based methods and natural bond orbital (NBO) interpretation. Both levels of theory showed that the axial conformations of compounds 13 are more stable than their equatorial conformations. The Gibbs free energy difference (G eqG ax) values (i.e. ΔG eq–ax) between the axial and equatorial conformations increase from compound 1 to compound 2 but decrease from compound 2 to compound 3. Based on the NBO results obtained, the AE associated with the electron delocalisation [i.e. Σ(endo-AEeq + exo-AEeq) ? Σ(endo-AEax + exo-AEax)] increase slightly from compound 1 to compound 2 but decrease from compound 2 to compound 3. Similar trend is also observed for the differences between the calculated total steric exchange energy values [i.e. Δ(TSEE)eq–ax]. On the other hand, the calculated differences between the dipole moment values of the axial and equatorial conformations [i.e. Δ(μeq–μax)] decrease from compound 1 to compound 3. These findings led to the proposal that the AE associated with the electron delocalisation (the hyperconjugation effect) is more significant for the explanation of the conformational preferences of compounds 13 than the electrostatic model. The correlations between the AE associated with the electron delocalisation, bond orders, TSEE, ΔG eq–ax, dipole–dipole interactions, structural parameters and conformational behaviours of compounds 13 have been investigated.  相似文献   

3.
The process of vision is initiated when the G protein-coupled receptor, rhodopsin (Rho), absorbs a photon and transitions to its activated Rho form. Rho binds the heterotrimeric G protein, transducin (Gt) inducing GDP to GTP exchange and Gt dissociation. Using nucleotide depletion and affinity chromatography, we trapped and purified the resulting nucleotide-free Rho·Gt complex. Quantitative SDS–PAGE suggested a 2:1 molar ratio of Rho to Gt in the complex and its mass determined by scanning transmission electron microscopy was 221 ± 12 kDa. A 21.6 Å structure was calculated from projections of negatively stained Rho·Gt complexes. The molecular envelope thus determined accommodated two Rho molecules together with one Gt heterotrimer, corroborating the heteropentameric structure of the Rho·Gt complex.  相似文献   

4.
O-Peracetylated or -perbenzoylated C-(1-bromo-1-deoxy-d-glycopyranosyl)formamides of d-gluco, d-galacto, and d-arabino configuration were reacted with Ag(I)-salts or HgO in nitrile solvents to give N-acyl-1-cyano-d-glycopyranosylamines with an axial C–N bond at the anomeric centre. In the presence of HgBr2, Hg(CN)2, or InCl3 the anomer of the above glycosylamine with an equatorial C–N bond was also isolated or detected. In CH3NO2 solutions as few as 5–10 equiv of the nitrile were sufficient to get acceptable yields for the products. Under similar conditions N-substituted C-(2,3,4,6-tetra-O-acetyl-1-bromo-1-deoxy-β-d-galactopyranosyl)formamides gave anomeric spiro-oxazoline derivatives which, upon mild acidic hydrolysis, opened up to di- and tripeptides of anomeric α-amino acids.  相似文献   

5.
Stereochemical properties of the glycosidic linkage have been studied by the quantum-chemical PCILO method, using 2-methoxytetrahydropyran as a model. Calculations of the two-dimensional, conformational (Φ, Ψ) maps showed that the rotation around the C-1---O-1 bond is more hindered than that around the O-1---C-6 bond, and that there are differences in the shape of the energy curve for the axial and equatorial forms of 2-methoxytetrahydropyran. The observed population of the five stable conformers at equilibrium (GG:GT:TG1:TG2:TT = 70.8:6.0:19.9:2.0:1.3) is consistent with the prediction of the anomeric and exo-anomeric effects. The calculated abundance (76.8%) of the axial form of 2-methoxytetrahydropyran is comparable with experimental results (77–80%) obtained by n.m.r. measurements in non-polar solvents. The energies found for individual conformers made it possible to calculate the magnitude of the anomeric effect (3 kJ/mol) and to determine, for the first time, the values of the exo-anomeric effect for axial (6 kJ/mol) and equatorial 2-methoxytetrahydropyran (7 kJ/mol). The calculated variations of the geometry arising from rotation around the C-1---O-1 bond are consistent with results obtained by statistical analysis of experimental data for - and β-glycosides. The results obtained, indicating that the energy, geometry, and electronic structure of glycosides are largely affected by the conformation of the acetal segment, are discussed from the point of view of conformational analysis of oligo- and poly-saccharides.  相似文献   

6.
7.
The first single-crystal X-ray diffraction study of tetrafluorinated monosaccharide derivatives is presented. Both α- and β-methyl 2,3-dideoxy-2,2,3,3-tetrafluoro-d-galactopyranoside anomers adopt the 4C1 conformation. The values for the C1–O1 and C1–O5 bond lengths and the O5–C1–O1–CH3 dihedral angles are in line with what can be expected from the anomeric and exo-anomeric effects. The chair conformations are slightly distorted, presumably due to repulsion between 1,3-diaxial C–O and C–F bonds. The asymmetric unit of both compounds contains up to three independent molecules, which differ in the conformation of the hydroxymethyl group (including in one case a ‘forbidden’ gg rotamer). The molecular packing of the β-anomer shows a clear segregation between fluorinated and hydrophilic domains, while for the α-anomer the regions of fluorine segregation are broken by interleafing of OMe groups. There is one close OH?F contact, which is likely to arise from the crystal packing. NMR studies show that the two anomers also adopt a 4C1 conformation in solution (D2O, CDCl3).  相似文献   

8.
Novel N-substituted-2-piperidones with a 1,4-benzodioxan ring were prepared and evaluated for their activity to induce lateral roots in lettuce seedlings. Compounds were obtained by aldol condensation of the lithium enolate of N-substituted-2-piperidones with 1,4-benzodioxan-6-carbaldehyde. Of the series compounds tested, N-cinnamyl-3-[1-(1,4-benzodioxan-6-yl)-1-hydroxymethyl]-2-piperidone (2e) had the highest activity. In seedlings treated with 10 ppm of 2e, all of the primary roots formed lateral roots. Only erythro-2e showed lateral root-inducing activity, while threo-2e was inactive.  相似文献   

9.
10.
A total of 16 pyrrolysine conformers in their zwitterionic forms are studied in gas and simulated aqueous phase using a polarizable continuum model (PCM). These conformers are selected on the basis of our study on the intrinsic conformational properties of non-ionic pyrrolysine molecule in gas phase [Das and Mandal (2013) J Mol Model 19:1695?1704]. In aqueous phase, the stable zwitterionic pyrrolysine conformers are characterized by full geometry optimization and vibrational frequency calculations using B3LYP/6-311++G(d,p) level of theory. Single point calculations are also carried out at MP2/6-311++G(d,p) level. Characteristic intramolecular hydrogen bonds present in each conformer, their relative energies, theoretically predicted vibrational spectra, rotational constants and dipole moments are systematically reported. The calculated relative energy range of the conformers at B3LYP/6-311++G(d,p) level is 5.19 kcal mol?1 whereas the same obtained by single point calculations at MP2/6-311++G(d,p) level is 4.58 kcal mol?1. A thorough analysis reveals that four types of intramolecular H-bonds are present in the conformers; all of which play key roles in determining the energetics and in imparting the observed conformations to the conformers. The vibrational frequencies are found to shift invariably toward the lower side of frequency scale corresponding to the presence of the H-bonds. This study also points out that conformers with diverse structural motifs may differ in their thermodynamical stability by a narrow range of relative energy. The effects of metal coordination on the relative stability order and structural features of the conformers are examined by complexing five zwitterionic conformers of pyrrolysine with Cu+2 through their carboxylate groups. The interaction enthalpies and Gibbs energies, rotational constants, vibrational frequencies and dipole moments of the metal complexes calculated at B3LYP level are also reported. The zwitterionic conformers of pyrrolysine are not stable in gas phase; after geometry optimization they are converted to the non-ionic forms.  相似文献   

11.
The molecular structure (hydrogen bonding, bond distances and angles), dipole moment and vibrational spectroscopic data [vibrational frequencies, IR and vibrational circular dichroism (VCD)] of cyclobutanone?HX (X?=?F, Cl) complexes were calculated using density functional theory (DFT) and second order Møller–Plesset perturbation theory (MP2) with basis sets ranging from 6–311G, 6–311G**, 6–311 + + G**. The theoretical results are discussed mainly in terms of comparisons with available experimental data. For geometric data, good agreement between theory and experiment is obtained for the MP2 and B3LYP levels with basis sets including diffuse functions. Surface potential energy calculations were carried out with scanning HCl and HF near the oxygen atom. The nonlinear hydrogen bonds of 1.81 Å and 175° for HCl and 1.71 Å and 161° for HF were calculated. In these complexes the C=O and H–X bonds participating in the hydrogen bond are elongated, while others bonds are compressed. The calculated vibrational spectra were interpreted and the band assignments reported are in excellent agreement with experimental IR spectra. The C=O stretching vibrational frequencies of the complexes show red shifts with respect to cyclobutanone.  相似文献   

12.
Vincent Madison 《Biopolymers》1977,16(12):2671-2692
A survey of over 50 crystal structures indicates that both imino acid and peptide derivatives of proline populate ring conformers consistent with the torsional potentials about single bonds. In both cases, lower barriers for rotation about C? N bonds relative to those about C? C bonds favor smaller values for dihedral angles about the former bonds. In peptides a minimum in the torsional potential about C? N bonds occurs at zero dihedral angle, further favoring small angles. The pyrrolidine-ring dihedral angles of the proline compounds in the solid state obey a cyclopentane-type pseudorotation function. Thus the puckering of the five-membered ring can be quantitatively described by two parameters. Consistent with small dihedral angles about C? N bonds, Cβ and/or Cγ are puckered out of the mean plane of the ring in nearly all of the nonstrained compounds. Utilizing the consistent force-field method of Lifson and coworkers [see A. Warshel, M. Levitt, and S. Lifson (1970) J. Mol. Spectrosc. 33 , 84] the intramolecular energy of five proline peptides was minimized with respect to all internal coordinates. In addition, the energy surface near minima was explored by constraining a particular dihedral angle and reminimizing the energy with respect to all remaining variables. In linear peptides two types of pyrrolidine-ring conformers have identical predicted energies. In the cyclic dipeptide cyclo (Pro-Gly) one of the ring conformers is favored by about 3 kcal/mol, while the cyclic tripeptide cyclo(Pro-Gly-Gly) favors the other conformer by a comparable margin. In agreement with observations in the solid state and in solution, Cβ and/or Cγ are puckered in the predicted conformers. A correlation between proline Φ and the details of the puckered conformation was predicted and found to match precisely conformers observed in crystals. For the diamides N-acetyl-L -proline-N′-methyl-amide and N-acetyl-L -proline-N′,N′-dimethylamide (AcProMe2A) 30% and 60% cis acetyl peptide bonds were predicted in good agreement with observations in nonpolar solvents for the respective compounds. The conformational distributions with respect to proline Ψ are also in accord with experimental observations. For AcProMe2A, a model for a -Pro-Pro-sequence in a peptide chain, this study is the first to predict stable conformers for proline Ψ either ca. ?50° or 140° for both cis and trans peptides.  相似文献   

13.
T Kozár  I Tvaroska 《Biopolymers》1990,29(12-13):1531-1539
The 2-methylaminotetrahydropyran was used as a model to study conformational properties of the N-glycosidic linkage in glycosylamines. Relaxed two-dimensional conformational (phi, psi) maps in 20 solvents were calculated by a method in which the total energy is divided into the energy of the isolated molecule and the solvation energy. Molecular geometry optimization has been carried out for each conformer using the quantum chemical method PCILO. The calculated variations of the geometry are consistent with the results obtained by the statistical analysis of available experimental data retrieved from the Cambridge Structural Database. The calculated abundances of conformers show that the polarity of the solvent has little effect on the anomeric ratio, and the form having the methylamino group equatorial is favored in all considered solvents.  相似文献   

14.
Some general rules governing hydrogen bonding at the ring oxygens of furanosides, pyranosides, and bridge oxygens of glycosides have been formulated from existing data on crystal structures of carbohydrates. Ring oxygens of the majority of the glycopyranosides in the hemiacetal or acetal form are involved in hydrogen bonding such that the hydrogen bond direction is usually equatorial to the ring plane and not axial. In contrast, there are no known examples of ring oxygens of glycofuranosides and methyl-glycopyranosides displaying hydrogen bonding in the crystal. Also, the bridge oxygens of glycosides are not involved in hydrogen bonding. The observed shortening in the exocyclic and endocyclic anomeric C(1)? O bonds and the geminal C? O bonds indicate that compounds with two oxygen atoms attached to the same saturated carbon atom may participate in double-bond-no-bond resonance interaction in the same manner as difluoromethane. It is also possible that under these circumstances the carbon atom exhibits greater than tetracovalency. The “anomeric effect” may also be related to (a) the differences in the “double bonding” or bond shortening in the anomeric C? O bonds of the anomeric glycopyranosides, (b) the shorter intramolecular O(1)…?O(5) non-bonded interaction, and (c) the smaller O(1)C(1)O(5) valence angle in the equatorial anomer compared to the axial anomer. An analysis has been made of the energetically preferred conformations about the glycosyl and glycosidic bonds of 1,4- and 1,3-polysuc-charides. In the 1a, 4e-glycopyranosides the projected angle ?1 [O(5)C(1)OR, where R = C or H] is positive, while it is negative in the 1e, 4e-glycopyranosides. Angle ?2 [C(1)OC(4′)C(3′)] is positive in both the 1,4-anomeric polyglycosides. 1e, 4e- and 1a, 4e -polysaccharides are stabilized by intramolecular O(5)…?H? O(3′) and O(2′)…?O(3′) hydrogen bonding, respectively, and generate linear and helical (cyclic) structures, respectively. 1e, 3e- and 1a, 3e-polysaccharides may be stablized by one of two possible intramolecular hydrogen-bonding schemes such that the 1a, 3e -polysaccharides generate helical structures while the 1a, 3e-polysaccharides generate nonhelical structures. The conformation about the C(5)? C(6) bond in the pyranosides falls into two groups where the angle ?00 [O(5)C(5)C(6)O(6)] is either positive, ~+60 ± 30°, or negative, ~–60 ± 30°, the former conformation being found more frequently. In the furanosides the latter conformation is preferred.  相似文献   

15.
16.
The rotational barriers ΔG (T) of the four atropisomeric polychlorinated biphenyls (PCBs) 2,2′,3,5′,6-pentachlorobiphenyl (PCB 95), 2,2′3,3′,4,6′-hexachlorobiphenyl (PCB 132), 2,2′,3,3′,6,6′-hexachlorobiphenyl (PCB 136), and 2,2′,3,4′,5′,6-hexachlorobiphenyl (PCB 149) were determined via on-line enantiomerization kinetics by a new stopped-flow multidimensional gas chromatographic technique (stopped-flow MDGC) employing Chirasil-Dex as chiral stationary phase for enantiomer separation. The calculated rotational barriers ΔG (T) of the trichloro-ortho-substituted atropisomers are 184 ± 2 kJ/mol for PCB 95, 189 ± 4 kJ/mol for PCB 132, and 184 ± 1 kJ/mol for PCB 149 at 300°C. The rotational barrier ΔG (T) of tetrachloro-ortho-substituted PCB 136 is at least (or higher than) 210 kJ/mol at 320°C. Chirality 10:316–320, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

17.
In the present study, the first quantum chemical calculations of structures and vibrational spectra of radicals of 1,4-naphthoquinone and 2-methoxy-1,4-benzoquinone that account for electron correlation are presented. In the case of 1,4-naphthoquinone a good agreement between calculated vibrational frequencies and 18O-shifts of the 1,4-naphthoquinone radical (protonated radical anion) with experimental data of a species detected after irradiation of vitamin K1 in solution is found. Our calculations, thus, support the previous assignment. In the case of 2-methoxy-1,4-benzoquinone we have localized the stable conformations with respect to the orientation of the methoxy group and we have determine the harmonic force fields for these structures. Our calculation suggest that, while the frequencies of the two conformers are similar, the 18O-shift of the most intensive absorptions in the spectral region between 1400 and 1700 cm–1 of the two conformers differ significantly and might serve as a tool to distinguish between the two conformers. The applied DFT method is shown to predict electron affinities which are systematically underestimated by 10%.  相似文献   

18.
The crystal structure and absolute configuration of the (?)589-dibenzoylmonohydrogentartrate salt of the cation [Co(pyridine)4Cl2]+ have been determined from a three-dimensional X-ray analysis. Single crystals were grown from dimethylsulfoxide: space group P212121, Z = 4, and cell dimensions a = 21.463(4), b = 23.112(3), and c = 7.490(1) Å. Full-matrix least-squares refinement on F converged at R = 0.075, 196 variables and 2029 observations. The cation has pseudotetragonal coordinate geometry, with axial Cl and equatorial N atoms. The dihedral angles between the pyridine ligands and the equatorial plane are 47(1), 39(1), 50(1), and 45(1)° and torsional isomerism is responsible for the solid-state chiroptical properties of the cation. The preferential crystallization of the P atropisomer of the cation is attributed to a general electrostatic attraction between cation and anion.  相似文献   

19.
Complexes of three Cd(II)-containing macrocyclic Schiff base complexes containing a phenanthroline ligand (L) of the type [CdLn(Cl)]+ (n=2,3,4), have been prepared via [1+1] cyclocondensation of 2,9-dicarboxaldehyde-1,10-phenanthroline and a number of linear triamines via a metal-templated reaction and coordination features have been examined. The ligands, L, are 16-, 17-, and 18-membered pentaaza macrocycles and all the complexes incorporate a 1,10-phenanthroline unit as an integral part of their cyclic structure. The complexes have been characterized by a variety of methods including IR, 1H, 13C, DEPT, COSY(H,H) and HMQC(H,C) NMR studies and MALDI mass spectrometry. The polymeric structure of was determined by X-ray crystallography, which showed that the complex cation consisted of a pentagonal bipyramidally coordinated Cd(II) ion. The seven-coordinated Cd(II) ion is ligated by the five nitrogen atoms of the macrocycle in the equatorial plane and has two bridging chloride ligands in the axial positions resulting in a ribbon of such complex ions. Supporting ab initio HF-MO calculations have been undertaken using the standard 3-21G and 6-31G basis sets.  相似文献   

20.
Among mammalian soft tissues, articular cartilage is particularly interesting because it can endure a lifetime of daily mechanical loading despite having minimal regenerative capacity. This remarkable resilience may be due to the depth-dependent mechanical properties, which have been shown to localize strain and energy dissipation. This paradigm proposes that these properties arise from the depth-dependent collagen fiber orientation. Nevertheless, this structure-function relationship has not yet been quantified. Here, we use confocal elastography, quantitative polarized light microscopy, and Fourier-transform infrared imaging to make same-sample measurements of the depth-dependent shear modulus, collagen fiber organization, and extracellular matrix concentration in neonatal bovine articular cartilage. We find weak correlations between the shear modulus |G| and both the collagen fiber orientation and polarization. We find a much stronger correlation between |G| and the concentration of collagen fibers. Interestingly, very small changes in collagen volume fraction vc lead to orders-of-magnitude changes in the modulus with |G| scaling as (vcv0)ξ. Such dependencies are observed in the rheology of other biopolymer networks whose structure exhibits rigidity percolation phase transitions. Along these lines, we propose that the collagen network in articular cartilage is near a percolation threshold that gives rise to these large mechanical variations and localization of strain at the tissue’s surface.  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