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 46 毫秒
1.
The partial isotope substitution for the change of geometrical parameters, interaction energies, and nuclear magnetic shielding tensors (σ) of dihydrogen‐bonded NH3X+···YBeH (X, Y = H, D, and T) systems is analyzed. Based on the theoretical calculation, the distance between heavy atoms RN···Be of NH3H+···DBeH is clearly found to be shorter than that in NH3D+···HBeH. Such apparently paradoxical geometrical isotope effect (GIE) on RN···Be is revealed by the cooperative effect of two kinds of (1) primary covalent‐bonded GIE and (2) secondary dihydrogen‐bonded one. We have demonstrated that (1) the covalent bond lengths become shorter by heavier isotope‐substitution and (2) the dihydrogen‐bonded distance RX···Y becomes shorter by heavier Y and lighter X isotope‐substitution due to the difference of electronic structure reflected by the nuclear distribution. We have also found that interaction energy of NH3H+···DBeH is stronger than that of NH3D+···HBeH and isotopic deshielding effect of magnetic shielding becomes large in lighter isotope. © 2013 Wiley Periodicals, Inc.  相似文献   

2.
Ab initio molecular orbital and DFT calculations have been carried out for three most stable dimers of parent nitrosamine (NA) in order to elucidate the structures and energetics of the dimers. The structures were optimized using HF, B3LYP, and MP2 methods with 6‐311+G(d,p) and 6‐311++G(2d,2p) basis sets. At the optimized geometries obtained at MP2/6‐311++G(2d,2p) level of theory, the energies were evaluated at QCISD/aug‐cc‐pVDZ and CCSD/aug‐cc‐pVDZ levels. The most stable dimer has two N? H···O?N hydrogen bonds and the least stable dimer has two N? H···N?O hydrogen bonds. The natural bond orbital analysis showed that the lpO(N) → BD*(N? N) and lpO(N) → BD*(N? Hb) interactions play a decisive role in the stabilization of the NH···O(N) hydrogen bonds in dimers. The atoms in molecules results reveal that the intermolecular N? H···O(N) H‐bonds in dimers have electrostatic character. © 2007 Wiley Periodicals, Inc. Int J Quantum Chem, 2008  相似文献   

3.
In this article, the geometry structures of hydrogen bond chains of formamide and N‐methylacetamide and their hydrogen‐bonded complexes with water were optimized at the MP2/6‐31G* level. Then, we performed Møller–Plesset perturbation method with 6‐311++g**, aug‐cc‐pvtz basis sets to study the cooperative influence to the total hydrogen bond energy by the N? H ··· OH2 and C?O ··· HOH hydrogen bonds. On the basis of our results, we found that the cooperativity of the hydrogen‐bonded complexes become weaker as N? H ··· OH2 and C?O ··· HOH hydrogen bonds replacing N? H ··· O?C hydrogen bonds in protein and peptide. It means that the N? H and C?O bonds in peptide prefer to form N? H ··· O?C hydrogen bond rather than to form C?O ··· HOH and N? H ··· OH2. It is significant for understanding the structures and properties of the helical or sheet structures of protein and peptide in biological systems. © 2011 Wiley Periodicals, Inc. Int J Quantum Chem, 2011  相似文献   

4.
The conformational study of β‐thioaminoacrolein was performed at various theoretical levels, HF, B3LYP, and MP2 with 6‐311++G(d,p) basis set, and the equilibrium conformations were determined. To have more reliable energies, the total energies of all conformers were recomputed at high‐level ab initio methods, G2MP2, G3, and CBS‐QB3. According to these calculations, the intramolecular hydrogen bond is accepted as the origin of conformational preference in thialamine (TAA) and thiolimine groups. The hydrogen bond strength in various resonance‐assisted hydrogen bond systems was evaluated by HB energy, geometrical parameters, topological parameters, and charge transfers corresponding to orbital interactions. Furthermore, our results reveal that the TAA tautomer has extra stability with respect to the other tautomers. The population analyses of the possible conformations by NBO predict that the origin of this preference is mainly due to the π‐electron delocalization in framework of TAA forms, especially usual πC?C → π*C?S and Lp (N) → π*C?C charge transfers. © 2010 Wiley Periodicals, Inc. Int J Quantum Chem, 2011  相似文献   

5.
In this article, the binding energies of 16 antiparallel and parallel β‐sheet models are estimated using the analytic potential energy function we proposed recently and the results are compared with those obtained from MP2, AMBER99, OPLSAA/L, and CHARMM27 calculations. The comparisons indicate that the analytic potential energy function can produce reasonable binding energies for β‐sheet models. Further comparisons suggest that the binding energy of the β‐sheet models might come mainly from dipole–dipole attractive and repulsive interactions and VDW interactions between the two strands. The dipole–dipole attractive and repulsive interactions are further obtained in this article. The total of N? H···H? N and C?O···O?C dipole–dipole repulsive interaction (the secondary electrostatic repulsive interaction) in the small ring of the antiparallel β‐sheet models is estimated to be about 6.0 kcal/mol. The individual N? H···O?C dipole–dipole attractive interaction is predicted to be ?6.2 ± 0.2 kcal/mol in the antiparallel β‐sheet models and ?5.2 ± 0.6 kcal/mol in the parallel β‐sheet models. The individual Cα? H···O?C attractive interaction is ?1.2 ± 0.2 kcal/mol in the antiparallel β‐sheet models and ?1.5 ± 0.2 kcal/mol in the parallel β‐sheet models. These values are important in understanding the interactions at protein–protein interfaces and developing a more accurate force field for peptides and proteins. © 2009 Wiley Periodicals, Inc. J Comput Chem 2010  相似文献   

6.
Theoretical calculations were performed to study the nature of the hydrogen bonds in the complexes HCHO···HSO, HCOOH···HSO, HCHO···HOO, and HCOOH···HOO. The geometric structures and vibrational frequencies of these four complexes at the MP2/6‐31G(d,p) and MP2/6‐311+G(d,p) levels are calculated by standard and counterpoise‐corrected methods, respectively. The results indicate that in the complexes HCHO···HSO and HCOOH···HSO the S? H bond is strongly contracted. In the S? H···O hydrogen bonds, the calculated blue shifts for the S? H stretching frequencies are in the vicinity of 50 cm?1. While in the complexes HCHO···HOO and HCOOH···HOO, the O? H bond is elongated and O? H···O red‐shifted hydrogen bonds are found. From the natural bond orbital analysis it can be seen that the X? H bond length in the X? H···Y hydrogen bond is controlled by a balance of four main factors in the opposite directions: hyperconjugation, electron density redistribution, rehybridization, and structural reorganization. Among them hyperconjugation has the effect of elongating the X? H bond. Electron density redistribution and rehybridization belong to the bond shortening effects, while structural reorganization has an uncertain influence on the X? H bond length. In the complexes HCHO···HSO and HCOOH···HSO, the shortening effects dominate which lead to the blue shift of the S? H stretching frequencies. In the complexes HCHO···HOO and HCOOH···HOO where elongating effects are dominant, the O? H···O hydrogen bonds are red‐shifted. © 2008 Wiley Periodicals, Inc. Int J Quantum Chem, 2009  相似文献   

7.
The reaction of singlet oxygen (1O2) generated by ultraviolet-A (UVA)-visible light (lambda > 330 nm) irradiation of air-saturated solutions of hematoporphyrin with phenolic compounds in the presence of a spin trap, 5,5-dimethyl-1-pyrroline-N-oxide (DMPO), gave an electron spin resonance (ESR) spectrum characteristic of the DMPO-hydroxyl radical spin adduct (DMPO-*OH). In contrast, the ESR signal of 5,5-dimethyl-2-pyrrolidone-N-oxyl, an oxidative product of DMPO, was observed in the absence of phenolic compounds. The ESR signal of DMPO-*OH decreased in the presence of either a *OH scavenger or a quencher of *O2 and under anaerobic conditions, whereas it increased depending on the concentration of DMPO. These results indicate both 1O2- and DMPO-mediated formation of free *OH during the reaction. When DMPO was replaced with 5-(diethoxyphosphoryl)-5-methyl-1-pyrroline-N-oxide (DEPMPO), no DEPMPO adduct of oxygen radical species was obtained. This suggests that 1O2, as an oxidizing agent, reacts little with DEPMPO, in which a strong electron-withdrawing phosphoryl group increases the oxidation potential of DEPMPO compared with DMPO. A linear correlation between the amounts of DMPO-*OH generated and the oxidation potentials of phenolic compounds was observed, suggesting that the electron-donating properties of phenolic compounds contribute to the appearance of *OH. These observations indicate that 1O2 reacts first with DMPO, and the resulting DMPO-1O2 intermediate is immediately decomposed/reduced to give *OH. Phenolic compounds would participate in this reaction as electron donors but would not contribute to the direct conversion of 1O2 to *OH. Furthermore, DEPMPO did not cause the spin-trapping agent-mediated generation of *OH like DMPO did.  相似文献   

8.
In line with previous work in which we established the factors that enhance attractive C? H···H? C dihydrogen interactions in alkanes, an extended theoretical analysis of noncovalent intermolecular interactions in group 14 hydrides is presented here. Remarkably, these weak interactions may play a major role in determining the crystal structures adopted by several families of molecules. A combined structural and computational analysis at the MP2 level allowed us to identify and characterize different interactions of the type E? H···H? E and E···H? E (E = Si, Ge, Sn, and Pb), and to find also the most suitable scenario for the establishment of each particular type. The nature of the interactions has been analyzed in terms of natural charges of the atoms involved and a topological analysis of the electron density of several dimers confirms the existence of H···H and H···E attractive contacts. We have observed that the interaction strength increases when descending down the periodic group and that silicon has a marked tendency to establish Si···H? Si interactions. A size‐dependent backbone effect that reinforces H···H dihydrogen interactions in polyhedral systems has also been found.  相似文献   

9.
The1 Hartree–Fock crystal orbital (CO) method in its linear combination of atomic orbitals form was applied to determine the band structure of histone proteins taking 0.041e charge transfer per nucleotide base from the PO groups of poly(guanilic acid) to the arginine, and lysine side chains in histones (see text). Assuming that there are infinite COs, perpendicular to the main chain, formed by the amide groups of one segment of the protein chain bound together by H‐bonds with the C?O groups of another segment of the chain, we have calculated the band structure. From this, we have determined the mobility using the deformation potential approximation. Multiplying this with the mobile electron concentration due to the charge transfer between the PO groups of DNA and the positive side chains in histones, we have obtained for the direct current (D.C.) electron conductivity σfib = 1.07 × 10?9 Ω?1 cm for a single fiber and after division by the cross‐section of 9.10 × 10?16 cm2, σspec = 1.18 × 106 Ω?1 cm?1 for the specific conductivity. © 2008 Wiley Periodicals, Inc. Int J Quantum Chem, 2009  相似文献   

10.
A novel single‐electron sodium bond system of H3C···Na? H (I), H3C···Na? OH(II), H3C···Na? F(III), H3C···Na‐CCH(IV), H3C···Na? CN (V) and H3C···Na? NC (VI) complexes has been studied by using MP2/6‐311++G** and MP2/aug‐cc‐pVTZ methods for the first time. We demonstrated that the single‐electron sodium bond H3C···Na? Y formed between H3C and Na? Y (Y?H, OH, F, CCH, CN, and NC) could induce the Na? Y increased and stretching frequencies of I–IV and VI are red‐shifted, including the Na? N bond in complex V is blue‐shifted abnormally. The interaction energies are calculated at two levels of theory [MP2, CCSD(T)] with different basis. The results shows that the strength of binding bond in group 2 (IV–VI) with π electrons are stronger than that of group 1 (I–III) without π electrons. For all complexes, the main orbital interactions between moieties H3C and Na? Y are LP1(C)→LP*1(Na). By comparisons with some related systems, it is concluded that the strength of single‐electron bond is increased in the order: hydrogen bond < bromine bond < sodium bond < lithium bond. © 2009 Wiley Periodicals, Inc. Int J Quantum Chem, 2011  相似文献   

11.
In this article, a new type of halogen‐bonded complex YCCX···HMY (X = Cl, Br; M = Be, Mg; Y = H, F, CH3) has been predicted and characterized at the MP2/aug‐cc‐pVTZ level. We named it as halogen‐hydride halogen bonding. In each YCCX···HMY complex, a halogen bond is formed between the positively charged X atom and the negatively charged H atom. This new kind of halogen bond has similar characteristics to the conventional halogen bond, such as the elongation of the C? X bond and the red shift of the C? X stretch frequency upon complexation. The interaction strength of this type of halogen bond is in a range of 3.34–10.52 kJ/mol, which is smaller than that of dihydrogen bond and conventional halogen bond. The nature of the electrostatic interaction in this type of halogen bond has also been unveiled by means of the natural bond orbital, atoms in molecules, and energy decomposition analyses. © 2009 Wiley Periodicals, Inc. J Comput Chem 2010  相似文献   

12.
To elucidate the nature of the Al? H···H? O dihydrogen bond and its effect on the reaction between diphenylmethanol and pyrazolate‐bridged dialuminum complex, a theoretical study was carried out using the ONIOM(B3LYP/6‐31+G(d,p):AM1) method. Calculations indicate that this reaction is a two‐step process. The first step is nucleophilic addition and the resulting intermediate is stabilized by an Al? H···H? O dihydrogen bond. Topology analyses based on the “atoms‐in‐molecules” theory show that the Al? H···H? O dihydrogen bond in dialuminum intermediate is stronger than normal hydrogen bond. This step is not barrierless, which is contrary to the result predicted by using simplified model. The second step, eliminating a molecule of dihydrogen, requires an activation free energy of 9.9 kcal/mol in gas phase, which implies the simplified model underestimates the energy barrier of this elimination step. ONIOM calculations also show that, using the simplified model without zero‐point energy correction, the dihydrogen bonding strength has been underestimated and unreliable results have been obtained. © 2007 Wiley Periodicals, Inc. Int J Quantum Chem, 2008  相似文献   

13.
The B24F224− anion, which was formed as a minor by‐product when the B12H122− anion was treated with F2 in liquid HF, has been isolated as its N(n‐Bu)4+ salt and characterized by 10B, 11B, and 19F NMR spectroscopy, electrospray mass spectrometry, cyclic voltammetry, single‐crystal X‐ray diffraction, and calculations at the DFT level of theory. The B24F224− anion has idealized D5 symmetry and consists of two B12F112− icosahedra linked by a 2c–2e boron–boron single bond with a B B distance of 1.725(4) Å. In the solid state, the anion interacts with eight N(n‐Bu)4+ cations via a network of 34 CH···FB hydrogen bonds with H· · ·F distances that range from 2.26 to 2.55 Å. These hydrogen bonds were successfully modeled by DFT calculations, which showed that the hydrogen bonds probably have a measurable, albeit subtle, effect on the structure of the B24F224−. © 2006 Wiley Periodicals, Inc. Heteroatom Chem 17:181–187, 2006; Published online in Wiley InterScience ( www.interscience.wiley.com ). DOI 10.1002/hc.20220  相似文献   

14.
15.
We tested correlations between lipophilicity parameters and the partitioning of sensitizers into membranes. For this purpose we investigated 17 porphyrins and two chlorins having various chemical structures. Some of these compounds possess an amphiphilic structure (including hematoporphyrin, deuteroporphyrin, mesoporphyrin, chlorin e6 and more). The others are very symmetrical sensitizers [meso-tetra(N-methyl-4-pyridyl)porphyrin, tetra-benzoporphyrin, coproporphyrin I dihydrochloride (CP), meso-tetra(4-carboxyphenyl)porphyrin (TCP) and meso-tetra(m-hydroxyphenyl)chlorin]. Our investigation also included two series of hematoporphyrins and protoporphyrins with varying lengths of alkylcarboxylate side groups. The partitioning of these compounds between the bulk aqueous phase and liposomes was studied by fluorescence methods, and a liposome-binding constant, Kb, was obtained. It was found that CP and TCP do not incorporate into the lipid phase at pH 7.3. An n-octanol-water partition coefficient (log P) and a distribution coefficient (log D) were predicted with a modeling software. The values of log D were also obtained experimentally. We found that for the studied molecules, the predicted log D correlated well with the measured values. The values of log D as well as log P, in turn, did not correlate nicely, for the whole group of studied compounds, with the binding constants to liposomes. However, in the case of porphyrins that share a similar structure, the Kb showed good linear correlation with both log P and log D. For the series of hematoporphyrins and protoporphyrins with different lengths of alkylcarboxyl groups, it was shown that prolongation of this group caused an increase in the lipophilicity and the liposome-binding constant. This effect is more pronounced for the proto- than for the hematoporphyrin series. The results highlight the possible use, as well as limitations, of lipophilicity parameters for the prediction of membrane binding.  相似文献   

16.
The excited-state intramolecular H-atom transfer of hypericin (Hyp) was investigated as a function of pH in monodispersed reverse micelles formed by sodium bis(2-ethylhexyl)sulfosuccinate/heptane/water and in complexes with Tb3+ under conditions in which one of the two carbonyl groups of Hyp is incapable of accepting a hydrogen atom. The results of pump-probe transient absorption experiments provide no evidence for a concerted H-atom transfer mechanism.  相似文献   

17.
Although ultraviolet-B (UV-B) radiation is potentially harmful, it is an important component of terrestrial radiation to which plants have been exposed since invading land. Since then, plants have evolved mechanisms to avoid and repair UV radiation damage; therefore, it is not surprising that photomorphogenic responses to UV-B are often assumed to be adaptations to harmful radiation. This presupposes that the function of the observed responses is to prevent UV damage. It has been hypothesized that, as with blue light, UV-B provides a signal important for normal plant development and might be perceived within developing plants through nondestructive processes, perhaps through UV-specific signal perception mechanisms. UV signal perception can lead to photomorphogenic responses that may confer adaptive advantages under conditions associated with high-light environments, such as water stress. Plant responses to UV radiation in this regard include changes in leaf area, leaf thickness, stomatal density, photosynthetic pigment production and altered stem elongation and branching patterns. Such responses may lead to altered transpiration rates and water-use efficiencies. For example, we found that the cumulative effect of ambient UV-B radiation upon stomatal density and conductance can lead to altered water-use efficiencies. In field settings, UV might more properly be viewed as a photomorphogenic signal than as a stressor. Hence, it might be insufficient to attempt to fully evaluate the adaptive roles of plant responses to UV-B cues upon stress tolerance by the simultaneous application of UV and drought stress during development. We propose that rather than examining a plant's response to combinations of stressors one might also examine how a plant's response to UV induces tolerance to subsequently applied stresses.  相似文献   

18.
The reaction of VI2 or TiI3, respectively, with ammonia in the presence of traces of water or oxygen, respectively, leads to [(NH3)5M? O? M(NH3)5]I4 · NH3 with M = V, Ti. Their structures were solved by X-ray single crystal data: Pbca (No. 61), Z = 4, M = V: a = 12.482(4) Å, b = 14.819(6) Å, c = 13.286(5) Å, N(F ? 3σF) = 983, N(variables) = 88, R/Rw = 0.053/0.063, M = Ti: a = 12.628(4) Å, b = 14.970(4) Å, c = 13.359(3) Å, N(F ? 3σF) = 1188, N(variables) = 88, R/Rw = 0.043/0.047. The structures consist of corner sharing octahedra double units [(NH3)5M? O? M(NH3)5]4+ with eclipsed conformation which are stacked together according to the motif of a distorted cubic face centered arrangement for the bridging oxygen atoms. IR spectroscopic investigations of the undeuterated vanadium compound and of 5% deuterated samples hint to N? H … I hydrogen bridge bonds and to remarkable π-bonding between the transition metal and the bridging oxygen atoms.  相似文献   

19.
The current cation framework representations are used to consider the organization principles associated with the crystal structures of a series of fluoridoborates with large cations and their topological analogues. The extra-framework anion components are differentiated into leading (template) and subordinate types with respect to their structure-directing function.  相似文献   

20.
Stannates Dy2Sn2O7 and Ho2Sn2O7 are produced by solid-phase synthesis from Dy2O3 (Ho2O3)–SnO2 stoichiometric mixtures by calcining at 1473 K. The molar heat capacity of holmium and dysprosium stannates is measured by differential scanning calorimetry (DSC) in the temperature range 370–1000 K. The experimental data are used to calculate thermodynamic properties (enthalpy change H°(T)–H°(370 K), entropy change S°(T)–S°(370 K), and the reduced Gibbs free energy Φ°(T)) of the synthesized compound.  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