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 703 毫秒
1.
The structures of eight symmetrically independent molecules of 2-methyl-2,4-pentanediol (MPD) in six crystal substances are studied based on the data retrieved from the Cambridge Structural Database (CSD). Coordinates of the most part of hydrogen atoms in MPD molecules were not determined experimentally or not presented in CSD, however, O...O distances provide the conclusion about the formation of intramolecular hydrogen bonds in four molecules. To perform quantum chemical calculations the absent hydrogen atoms were added. The choice of H atomic positions in hydroxyl groups are based on the analysis of possible formation of intra- and intermolecular hydrogen bonds by MPD molecules in the respective crystals. The DFT method with the B3PW91 functional and the 6-31G(d,p) basis set is used to carry out for the first time: 1) the calculation of dipole moments and energies for MPD molecules in “crystal” conformations; 2) the optimization of the structure of these molecules with the calculation of dipole moments for the conformations corresponding to the local energy minima. It is found that among the molecules with the experimental geometric parameters one of the conformations without intramolecular hydrogen bonds is most favorable (μ = 0.56 D). As a result of the energy minimization of eight “crystal” conformations in vacuum, five energetically different conformers are obtained. Among them the conformer with the intramolecular hydrogen bond has the lowest energy (μ = 3.53 D). Four variants of the molecular structure correspond to it in the considered crystals, out of which two are R-enantiomers and two S-enantiomers.  相似文献   

2.
The effects of three dihydric phenols on the thermal properties of poly‐(ε‐caprolactone) (PCL) were investigated by DSC. The thermal properties of PCL were found to be greatly modified by the addition of 4,4′‐dihydroxydiphenyl ether (DHDPE). When the content of DHDPE reached 40%, PCL that was a semicrystalline polymer in the pure state changed to a fully amorphous elastomer. Fourier transform infrared (FTIR) spectroscopy was also applied to investigate the specific interaction between PCL and DHDPE. The formations of intermolecular hydrogen bonds between the carbonyl groups of PCL and the hydroxyl groups of DHDPE were discovered. By applying the Beer–Lambert law and a curve‐fitting program, the fractions of hydrogen‐bonded carbonyl groups were quantitatively analyzed. Although one DHDPE molecule had the potentiality to form two hydrogen bonds with PCL chains, the values of the fraction of the hydroxyl group involved in the intermolecular hydrogen bond were so little that from a statistical point of view, the formation of two hydrogen bonds was very difficult for every DHDPE molecule. Both DSC and FTIR revealed that 4,4′‐dihydroxydiphenyl methane and 4,4′‐dihydroxyphenyl had the ability to form hydrogen bonds with PCL, which were strongly affected by the polarity of the group linking two hydroxyphenyls and the flexibility of the molecular chain. The stronger the polarity of the group and the better the flexibility of molecular chain, the more tendencies dihydric phenol had to form intermolecular hydrogen bonds with PCL. © 2001 John Wiley & Sons, Inc. J Polym Sci Part B: Polym Phys 39: 2108–2117, 2001  相似文献   

3.
The conformational behavior of cellobiose was studied by molecular dynamics simulation in a periodic box of waters. Several different initial conformations were used and the results compared with equivalent vacuum simulations. The average positions and rms fluctuations within single torsional conformations of cellobiose were affected only slightly by the solvent. However, water damped local torsional librations and transitions. The conformational energies of the solute and their fluctuations were also sensitive to the presence of solvent. Intramolecular hydrogen bonding was weakened relative to that observed in vacuo due to competition with solvating waters. All cellobiose hydroxyl groups participated in intermolecular hydrogen bonds with water, with approximately eight hydrogen bonds formed per glucose ring. The hydrogen bonding was predominantly between water hydrogens and solute hydroxyl oxygens. Intermolecular hydrogen bonding to ring and bridge oxygens was seldom present. The diffusion coefficients of both water and solute agree closely with experimental values. Water interchanged rapidly between the solvating first shell and the bulk on the picosecond time scale. © 1993 John Wiley & Sons, Inc.  相似文献   

4.
In the title compound, C31H40N2O·H2O, the outer two six‐membered rings are in chair conformations, while the central ring is in an 8β,9α‐half‐chair conformation. The five‐membered ring adopts a 13β‐envelope conformation and the cyano­benzyl­idene moiety has an E configuration with respect to the hydroxyl group at position 17. The steroid nuclei are linked by intermolecular O—H?O and O—H?N hydrogen bonds to form a molecular network. The molecular packing has an interesting feature, with the steroids aligned parallel to the b axis, forming a closed loop through hydrogen bonds linked via water mol­ecules.  相似文献   

5.
Density functional calculations of optimized geometries for the migration of single hydrogen and hydroxyl groups on graphene are performed. It is shown that the migration energy barrier for the hydroxyl group is three times larger than for hydrogen. The crucial role of supercell size for the values of the migration barriers is discussed. The paired migration of hydrogen and hydroxyl groups has also been examined. It could be concluded that hydroxyl group based magnetism is rather stable in contrast with unstable hydrogen based magnetism of functionalized graphene. The role of water in the migration of hydroxyl groups is also discussed, with the results of the calculations predicting that the presence of water weakens the covalent bonds and makes these groups more fluid. Increasing the number of water molecules associated with hydroxyl groups provides an increase of the migration energy.  相似文献   

6.
Ro 31-8959 is a highly potent inhibitor of HIV-1 proteinase in phase III clinical trials for treatment of AIDS. It is also the first subnanomolar inhibitor that demonstrated reversed stereochemical preference at the central hydroxyl group. Free energy perturbation calculations have been carried out to rationalize the preference for the R-diastereomer by consideration of two models of the (weaker) S-diastereomer. In the first model, the central hydroxyl group makes only one hydrogen bond with the active site aspartates, whereas the hydroxyl group in the second model makes at least three strong hydrogen bonds. Using the first model, the free energy difference in binding of Ro 31-8959 and its S-diastereomer is calculated to be 3.4 kcal/mol, which is in close agreement with the experimental value. Although the second model has a more favorable interaction with the active site aspartates compared to the first model, it has a higher energy N-axial conformation at the decahydroisoquinoline group in P. We show here that the two contributions cancel each other and the two models of S-diastereomer are predicted to have equivalent binding. The stereochemical preference in a hydroxyethylamine series of inhibitors appears to be affected by both intermolecular and intramolecular (conformational) energies. The binding data on the proline containing inhibitors are rationalized based on these results. © 1994 by John Wiley & Sons, Inc.  相似文献   

7.
The crystal structure of a novel non-peptidic HIV-1 protease inhibitor derived by simple solid-state dimerization of 4-aryl-1,4-dihydropyridines, reveals a strained central cage and the conformation of its phenyl, benzyl, and hydroxymethylene substituents. The polycyclic cage includes two nearly flat cyclobutane rings and four fused piperidine rings in boat conformations. The cage geometry reveals two unexpected features, namely marked distortions of the valence angles in every second piperidine and a shortening of one of the cyclobutane bonds. The molecule displays exact centrosymmetry, but the central cage and the hydroxymethylene substituents also approximate the C2-symmetry of the target enzyme. The two independent hydroxyl groups are involved in intermolecular hydrogen bonding, one as a donor, the other as an acceptor. The disposition of the hydroxyl groups in the molecular framework is compatible with the dual role of the inhibitor in the active-site cavity of HIV-1 protease, whereby one OH group is hydrogen-bonded to the catalytic aspartates, whereas another one provides an interface to the locked flaps of the enzyme.  相似文献   

8.
As part of the series investigating the structural features of C-terminal amidated amino acids and peptides, three crystal structures of Z-Gly-Phe-NH2, Tyr-Lys-NH2, and Asp-Phe-NH2 were analyzed by the X-ray diffraction method, and their molecular conformations and intermolecular interactions were investigated. Although the respective dipeptides exhibited an energetically allowable torsion angle concerning each backbone or side chain, the observed extended (Z-Gly-Phe-NH2, Asp-Phe-NH2) and folded (Tyr-Lys-NH2) conformations were considerably different from those of the corresponding unamidated peptides, due to the conformational flexibility of the respective dipeptides. The comparison between the crystal packings of the amidated and unamidated dipeptides indicated that the C-terminal amides tend to associate with the same neighboring group through hydrogen bonds, in which both the amide NH and O=C groups participate, while the unamidated peptides prefer a linear molecular connection, where both or either of the two carboxyl oxygens participate in the hydrogen bond formation. The difference in hydrogen bonding ability between the C-terminal amide and carboxyl groups has been considered to be based on the structural data of the related peptides analyzed so far.  相似文献   

9.
In the title compound, C30H50O3, the three six‐membered rings adopt chair, twist and twist‐boat conformations. The five‐membered ring is in a slightly distorted envelope conformation. The substituent on the five‐membered ring is in an extended conformation, with its two hydroxyl O atoms forming an intramolecular hydrogen bond. One of these O atoms also forms an intermolecular hydrogen bond with the oxy­gen of the carbonyl group in a neighbouring mol­ecule.  相似文献   

10.
The antisymmetry of proton configurations was studied for the hexagonal water rings with different conformations. The change in the direction of all hydrogen bonds was used as an additional symmetry operation. The ring configuration energies were calculated using the intermolecular interaction potentials. For different ring conformations, the relationships between antisymmetry and energy were analyzed and compared.  相似文献   

11.
The conformations of 2-cyclopenten-1-ol (2CPOL) have been investigated by high-level theoretical computations and infrared spectroscopy. The six conformational minima correspond to specific values of the ring-puckering and OH internal rotation coordinates. The conformation with the lowest energy possesses intramolecular π-type hydrogen bonding. A second conformer with weaker hydrogen bonding has somewhat higher energy. Ab initio coupled-cluster theory with single and double excitations (CCSD) was used with the cc-pVTZ (triple-ζ) basis set to calculate the two-dimensional potential energy surface (PES) governing the conformational dynamics along the ring-puckering and internal rotation coordinates. The two conformers with the hydrogen bonding lie about 300 cm−1 (0.8 kcal/mole) lower in energy than the other four conformers. The lowest energy conformation has a calculated distance of 2.68 Å from the hydrogen atom on the OH group to the middle of the C=C double bond. For the other conformers, this distance is at least 0.3 Å longer. The infrared spectrum in the O-H stretching region agrees well with the predicted frequency differences between the conformers and shows the conformers with the hydrogen bonding to have the lowest values. The infrared spectra in other regions arise mostly from the two hydrogen-bonded species.  相似文献   

12.
Isomerism, conformations, and molecular structure of a model molecule of vitamin K1 with a truncated side chain have been studied by the density functional theory calculations using B3LYP method and double- and triple-ζ correlation consistent basis sets. The conformations of two possible (E and Z) isomers, formed by the rotations around three single C–C bonds closest to the naphthoquinone ring, have been studied. The lowest energy conformers are stabilized by additional hydrogen bonds between hydrogen atoms of the side chain and an oxygen atom in the naphthoquinone subunit. It is interesting to note that the structure of the energetically preferred conformer of the E-isomer (3c) has been found to be similar to the solid state structures of phylloquinones in the photosystem I of cyanobacterium Synechococcus elongatus. The excited electronic states of two lowest energy conformers have also been investigated.  相似文献   

13.
We discovered that poly(vinylphenol) (PVPh) possesses an extremely low surface energy (15.7 mJ/m2) after a simple thermal treatment procedure, even lower than that of poly(tetrafluoroethylene) (22.0 mJ/m2) calculated on the basis of the two-liquid geometric method. Infrared analyses indicate that the intermolecular hydrogen bonding of PVPh decreases by converting the hydroxyl group into a free hydroxyl and increasing intramolecular hydrogen bonding after thermal treatment. PVPh results in a lower surface energy because of the decrease of intermolecular hydrogen bonding between hydroxyl groups. In addition, we also compared surface energies of PVPh-co-PS (polystyrene) copolymers (random and block) and their corresponding blends. Again, these random copolymers possess a lower fraction of intermolecular hydrogen bonding and surface energy than the corresponding block copolymers or blends after similar thermal treatment. This finding provides a unique and easy method to prepare a low-surface-energy material through a simple thermal treatment procedure without using fluoro polymers or silicones.  相似文献   

14.
Three hydrogen bonding complexes of the gauche‐1PA dimer (GG), trans‐1PA dimer (TT) and mixed dimer (GT) have been calculated for the geometry conformations and excited‐state energies. The electron distribution at the site of C‐O of H‐donor moiety in HOMO transfers to the direction of O‐H of H‐acceptor moiety in LUMO. The hydrogen bond between two 1PAs is the bridge of the intermolecular charge transfer. By the Zhao and Han's excited‐state hydrogen bonding dynamics rule, the first excited‐state hydrogen bonding change has been discussed without optimizing the excited‐state geometry conformations. According to the distinct difference between GT and GG (TT), we concluded that two gauche‐1PA monomers of one dimer are transformed at the same time to two trans‐1PA monomers.  相似文献   

15.
Robert Bau  Irina Tsyba 《Tetrahedron》1999,55(52):2871-14846
Amikacin is one of the important aminoglycoside antibiotics used against gram-negative bacteria. Here we report the crystal structure of amikacin that has been crystallized by vapor diffusion against polyethylene glycol. The molecule exists in a long, extended conformation, with all three six-membered rings in chair conformations and connected together by -glycosidic linkages. The orientation between the A, B and C rings of the molecule is maintained by intramolecular hydrogen bonds involving the O5 hydroxyl group and the amide NH group.  相似文献   

16.
17.
A flexible protein-peptide docking method has been designed to consider not only ligand flexibility but also the flexibility of the protein. The method is based on a Monte Carlo annealing process. Simulations with a distance root-mean-square (dRMS) virtual energy function revealed that the flexibility of protein side chains was as important as ligand flexibility for successful protein-peptide docking. On the basis of mean field theory, a transferable potential was designed to evaluate distance-dependent protein-ligand interactions and atomic solvation energies. The potential parameters were developed using a self-consistent process based on only 10 known complex structures. The effectiveness of each intermediate potential was judged on the basis of a Z score, approximating the gap between the energy of the native complex and the average energy of a decoy set. The Z score was determined using experimentally determined native structures and decoys generated by docking with the intermediate potentials. Using 6600 generated decoys and the Z score optimization criterion proposed in this work, the developed potential yielded an acceptable correlation of R(2) = 0.77, with binding free energies determined for known MHC I complexes (Class I Major Histocompatibility protein HLA-A(*)0201) which were not present in the training set. Test docking on 25 complexes further revealed a significant correlation between energy and dRMS, important for identifying native-like conformations. The near-native structures always belonged to one of the conformational classes with lower predicted binding energy. The lowest energy docked conformations are generally associated with near-native conformations, less than 3.0 Angstrom dRMS (and in many cases less than 1.0 Angstrom) from the experimentally determined structures.  相似文献   

18.
Microsolvated formamide clusters have been generated in a supersonic jet expansion and characterized using Fourier transform microwave spectroscopy. Three conformers of the monohydrated cluster and one of the dihydrated complex have been observed. Seven monosubstituted isotopic species have been measured for the most stable conformer of formamide...H(2)O, which adopts a closed planar ring structure stabilized by two intermolecular hydrogen bonds (N-H...O(H)-H...O=C). The two higher energy forms of formamide...H(2)O have been observed for the first time. The second most stable conformer is stabilized by a O-H...O=C and a weak C-H...O hydrogen bond, while, in the less stable form, water accepts a hydrogen bond from the anti hydrogen of the amino group. For formamide...(H(2)O)(2), the parent and nine monosubstituted isotopic species have been observed. In this cluster the two water molecules close a cycle with the amide group through three intermolecular hydrogen bonds (N-H...O(H)-H...O(H)-H...O=C), the nonbonded hydrogen atoms of water adopting an up-down configuration. Substitution (r(s)) and effective (r(0)) structures have been determined for formamide, the most stable form of formamide...H(2)O and formamide...(H(2)O)(2). The results on monohydrated formamide clusters can help to explain the observed preferences of bound water in proteins. Clear evidence of sigma-bond cooperativity effects emerges when comparing the structures of the mono- and dihydrated formamide clusters. No detectable structural changes due to pi-bond cooperativity are observed on formamide upon hydration.  相似文献   

19.
Ab initio geometry optimization was carried out on 10 selected conformations of maltose and two 2‐methoxytetrahydropyran conformations using the density functional denoted B3LYP combined with two basis sets. The 6‐31G* and 6‐311++G** basis sets make up the B3LYP/6‐31G* and B3LYP/6‐311++G** procedures. Internal coordinates were fully relaxed, and structures were gradient optimized at both levels of theory. Ten conformations were studied at the B3LYP/6‐31G* level, and five of these were continued with full gradient optimization at the B3LYP/6‐311++G** level of theory. The details of the ab initio optimized geometries are presented here, with particular attention given to the positions of the atoms around the anomeric center and the effect of the particular anomer and hydrogen bonding pattern on the maltose ring structures and relative conformational energies. The size and complexity of the hydrogen‐bonding network prevented a rigorous search of conformational space by ab initio calculations. However, using empirical force fields, low‐energy conformers of maltose were found that were subsequently gradient optimized at the two ab initio levels of theory. Three classes of conformations were studied, as defined by the clockwise or counterclockwise direction of the hydroxyl groups, or a flipped conformer in which the ψ‐dihedral is rotated by ∼180°. Different combinations of ω side‐chain rotations gave energy differences of more than 6 kcal/mol above the lowest energy structure found. The lowest energy structures bear remarkably close resemblance to the neutron and X‐ray diffraction crystal structures. © 2000 John Wiley & Sons, Inc. * J Comput Chem 21: 1204–1219, 2000  相似文献   

20.
The isotope effect in the melting temperature of ice Ih has been studied by free energy calculations within the path integral formulation of statistical mechanics. Free energy differences between isotopes are related to the dependence of their kinetic energy on the isotope mass. The water simulations were performed by using the q-TIP4P/F model, a point charge empirical potential that includes molecular flexibility and anharmonicity in the OH stretch of the water molecule. The reported melting temperature at ambient pressure of this model (T=251?K) increases by 6.5±0.5 and 8.2±0.5?K upon isotopic substitution of hydrogen by deuterium and tritium, respectively. These temperature shifts are larger than the experimental ones (3.8 and 4.5 K, respectively). In the classical limit, the melting temperature is nearly the same as that for tritiated ice. This unexpected behavior is rationalized by the coupling between intermolecular interactions and molecular flexibility. This coupling makes the kinetic energy of the OH stretching modes larger in the liquid than in the solid phase. However, the opposite behavior is found for intramolecular modes, which display larger kinetic energy in ice than in liquid water.  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