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 effect of molecular interaction on the O? H stretching force constant of methanol (MeOH) is reported for its associated species. The various electron donors (D) and acceptors (A) considered include organic molecules such as methanol, dimethylether, acetone, acetonitrile, dimethyl formamide, pyridine, and ions such as F?, Cl?, Li+, and H+. The variation in the O? H stretching force constant of MeO? H…?D species on interaction with the electron acceptor such as in the species is explained on the basis of the cooperativity effect. (CE ). The effect is discussed in terms of the relationship CE = (ΔF/F) × 100, where ΔF is the reduction is force constant of the hydrogen-bonded O? H stretching mode of the associated methanol species MeOH…?D when the lone pair electrons on oxygen of the methanol molecule are involved in hydrogen bonding with A, and F is the hydrogen-bonded O? H stretching force constant of the species when the lone pair electrons are free. The cooperativity effect (CE ) is found to increase with increasing electron acceptor and electron donor capacities of A and D. The calculated force constants are compared with the experimental results.  相似文献   

2.
Stretching force constants for formamide and its seven associated species involving two to four molecules hydrogen-bonded through linear and cyclic configurations and 10 structures containing formamide hydrogen-bonded with one to five water molecules are reported. Since ab initio calculations are rather inconvenient to perform on such big clusters and are time-consuming, CNINDO MO calculations were carried out using the gradient method. The results demonstrate, on the one hand, the feasibility of semiempirical calculations for the evaluation of trends in force constants for big clusters where generally ab initio calculations become much involved and, on the other hand, explain the effect of hydrogen bonding and cooperativity on force constants and vibrational spectra of biologically important systems composed of formamide in the condensed phase and its aqueous solutions. The C?O and N? H stretching force constants are found to reduce significantly on hydrogen bonding. The reduction in force constant is further enhanced when two cyclic dimers become associated through a linear hydrogen bond. The results indicate justification for the stabilization of the formamide structure with two cyclic dimers hydrogen-bonded together. The reduction in the force constants on hydrogen bonding also reflect the cooperativity contribution. The C?O and C? N stretching force constants for the structures corresponding to formamide in liquid and aqueous solution phases are in agreement with the experimental vibrational frequencies reported.  相似文献   

3.
CNDO/Force calculations have been performed on a series of molecules, H2CO, F2CO, CF4, CHF3, CH2F2 and CH3F. The optimum geometries and force fields are reported. It is found that the method can successfully predict the geometries of polyatomic molecules. The bending force constants and interaction force constants are, in general, comparable with experimental values both with respect to sign and magnitude. The stretching force constants have higher values than the experimental force constants. However, the trend in stretching force constants of a series of molecules is comparable with that of the corresponding experimental values.  相似文献   

4.
The stretching force constant correlates satisfactorily with theWiberg's bond order,K(N-O)=11.791 [p(N-O)]–9.565, with a correlation coefficient of 0.958, for nitrogen—oxygen systems.
Empirische Korrelation zwischen der Kraftkonstante der Streckschwingung und der Bindungsordnung in Stickstoff-Sauerstoff-Systemen (Kurze Mitteilung)
Zusammenfassung Die KraftkonstanteK der Streckschwingung läßt sich in befriedigender Weise mit derWibergschen Bindungsordnungp korrelieren:K(N-O)=11.791 [p(N-O)]-9.565. Für Stickstoff-Sauerstoff-Systeme beträgt der Korrelationskoeffizient 0.958.
  相似文献   

5.
In the present work we propose a numerical approach to estimate the harmonic and anharmonic force constant matrices, supposing we are able to compute analytically the first order derivative vector of the potential energy surface with respect to the internal coordinates. We use a polynomial least square fit to interpolate this gradient in the stationary point region. The structure of the regression matrix shows that the harmonic force constant matrix may be obtained even for large molecules; the evaluation of the anharmonic contributions request slightly more labor but is possible for 5 to 7 atoms. The present work is applicable even at the CI level and the number of computations remains small. We use the experimental planification to select the geometries to be computed in order to improve the estimation of the regression coefficients i.e. this means to lower their variance.chercheur qualifié du Fonds National Belge de la Recherche Scientifique.  相似文献   

6.
CNDO/force calculations on the complexes of nitromethane and acetonitrile with H+, Li+, F and Cl have been carried out to determine their geometrical parameters and stretching force constants. The results are discussed in terms of specific interactions between the ions and the molecules.  相似文献   

7.
The five trimers of H2O···HNC···H2O, H2O···H2O···HNC, HNC···H2O···H2O, H2O···HNC···HNC, and HNC···HNC···H2O have been studied with quantum chemical calculations. Their structures, harmonic vibrational frequencies and interaction energies have been calculated at the B3LYP and MP2 levels with the aug‐cc‐pVDZ and aug‐cc‐pVTZ basis sets. The cooperative effect on these properties has also been studied quantitatively. For HNC:(H2O)2 systems, the cyclic H2O···H2O···HNC trimer is most stable with an interaction energy of ?16.01 kcal/mol and a large cooperative energy of ?3.25 kcal/mol at the MP2/aug‐cc‐pVTZ level. For H2O:(HNC)2 systems, the interaction energy and cooperative energy in the H2O···HNC···HNC trimer are larger than those in the HNC···HNC···H2O trimer. The NH stretch frequency has a blue shift for the terminal HNC molecule in the HNC···H2O···H2O and HNC···HNC···H2O trimers and a red shift in other cases. A many‐body analysis has also been performed to understand the interaction energies in these hydrogen‐bonded clusters. © 2010 Wiley Periodicals, Inc. Int J Quantum Chem, 2011  相似文献   

8.
In the present article, two focal subjects, i.e., hydration of the NO and associated ion species in the Mg(NO3)2 solution are researched by using the ab initio method. Nitrate ions with the hydration number of 1–6 are optimized at the HF/6‐31+G* level. Their relative energies, binding energies, and v1‐NO frequencies are also presented. The investigation of the binding energies shows the hydration number is 3–6 in the solvent abundant environment. The associated species, including ion pairings, triple‐ and multiple‐ion clusters, are also optimized at the same level and their v1‐NO frequencies are calculated for comparing with the results in experiments. From the comparison, the new associated process via aqueous free ions → solvent‐shared ion pairings → solvent‐shared triple and multiple ion clusters → contact multiple ion clusters → amorphous crystal is proposed. © 2010 Wiley Periodicals, Inc. J Comput Chem 2010  相似文献   

9.
The change in the electric field at a nucleus in a molecule due to bond stretch is related to the force constant of the stretched bond. The validity of this relationship using approximate wave functions at the SCF and MP2 levels of theory is tested for the diatomic molecules H2, HF, CO, and N2. The effect of basis set variation on H2 is also investigated. © 1997 John Wiley & Sons, Inc. J Comput Chem 18 : 1664–1667, 1997  相似文献   

10.
The recently introduced adaptive force matching (AFM) method is used to develop a significantly improved pair‐wise nonpolarizable potential for water. A rigid version of the potential is also presented to enable larger time steps for biological simulations. In this work, it is demonstrated that the AFM method can be used to systematically assess the importance of each functional term during the construction of a force field. For a water potential, it is established that a single off‐atom charge center (M) in the plane of water outperforms two out‐of‐plane charge sites for reproducing intermolecular forces. The four‐site pair‐wise nonpolarizable force field developed in this work rivals some of the most sophisticated polarizable models in terms of reproducing accurate ab initio forces. The force fields are parameterized to perform best in the temperature range from 0 to 40°C. Equilibrium and dynamical properties calculated with the flexible and rigid force fields are in good agreement with experimental results. For the flexible model, the agreement improves when path integral simulation is performed. These force fields provide high‐quality results at a very low computational cost and are thus well suited to atomistic scale biological simulations. The AFM method provides a mechanism for selecting important terms in force field expressions and is a very promising tool for producing accurate force fields in condensed phases. © 2010 Wiley Periodicals, Inc. J Comput Chem, 2011  相似文献   

11.
Correlations are established between selected substituent parameters (σIP, σRPPORO) and the carbonyl stretching force constant, k(CO), for 28 mono and poly-substituted tricarbonylchromium-complexed arene compounds. On the basis of the statistical results it is concluded that the overall electronic substituent effect transmitted to the carbonyl groups involves both mesomeric and inductive mechanisms. Within the restricted domain, including substituent group and benzene ring, transmission proceeds largely by resonance, with a minor inductive (through-bond and field) effect operative in the same domain. Further transmission from the substituted arene ring to the chromium atom predominantly involves an inductive mechanism. This result, in support of existing literature data, suggests appreciable participation of the ring carbon σ framework in the metal—ring bond formation.  相似文献   

12.
Density functional theory at the B3LYP level with the 6‐311G** basis set is performed to calculate the systems consisting of up to four hydrazoic acid molecules. The dimers are found to exhibit cyclic and chain structures with N … H contacts at ca. 2.1–2.7 Å. However, there are only cyclic structures with N … H contacts at ca. 2.0–2.3 Å and 2.0–2.1 Å in the trimer and tetramer, respectively. Hydrogen bond distances in the trimer and tetramer are shorter than those in the cyclic dimer as a result of the stronger interaction between molecules. The contribution of cooperative effect to the interaction energy is significant. After the correction of the basis set superposition error and zero‐point energy, the binding energies are ?10.69, ?29.34, and ?54.26 kJ·mol?1 for the most stable dimer, trimer, and tetramer, respectively. The calculated IR shifts for N? H stretching mode increase with the size of the cluster growths, reaching more than 200 cm?1 in the tetramer. For the most stable clusters, the transition from the monomer to dimer, dimer to trimer, and trimer to tetramer involve changes of ?14.40, ?25.68, and ?31.88 kJ·mol?1 for the enthalpies at 298.15 K and 1atm, respectively. We also perform Mulliken populations analysis and find the Mulliken populations on intermolecular N … H increasing in the sequence of the dimer, trimer, and tetramer. © 2003 Wiley Periodicals, Inc. Int J Quantum Chem 94: 279–286, 2003  相似文献   

13.
The shape of micro‐droplets of water on a pure copper surface was investigated using the a.c. non‐contact mode of an atomic force microscope (AFM) by applying different attractive forces between the cantilever tip and the liquid surface. The forces largely influenced the observed radii of micro‐droplets; the influence can be reduced significantly by reducing the force. The same attractive force between the cantilever tip and the micro‐droplets is necessary when comparing the contact angles of micro‐droplets on different surfaces. Furthermore, the values of the contact angles of the micro‐droplets should be the average of those on at least four sides of the droplets. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

14.
通过典型的双分子亲核取代和烯烃亲电加成反应,讨论了反应过程中,化学键伸缩振动力常数和分子中1个电子所受到的作用势(D_(pb))之间的关联.研究表明分子中1个电子所受到的作用势能够很好地描述反应中键形成和断裂时的化学键强度.  相似文献   

15.
Interactions of dimethyl sulfoxide with carbon dioxide and water molecules which induce 18 significantly stable complexes are thoroughly investigated. An addition of CO2 or H2O molecules into the DMSO⋯1CO2 and DMSO⋯1H2O systems leads to an increase in the stability of the resulting complexes, in which it is larger for a H2O addition than a CO2. The overall stabilization energy of the DMSO⋯1,2CO2 is mainly contributed by the S=O⋯C Lewis acid–base interaction, whereas the O − H⋯O hydrogen bond plays a significant role in stabilizing complexes of DMSO⋯1,2H2O and DMSO⋯1CO2⋯1H2O. Remarkably, the complexes of DMSO⋯2H2O are found to be more stable than DMSO⋯1CO2⋯1H2O and DMSO⋯2CO2. The level of the cooperativity of multiple interactions in ternary complexes tends to decrease in going from DMSO⋯2H2O to DMSO⋯1CO2⋯1H2O and finally to DMSO⋯2CO2. It is generally found that the red shift of the O − H bond involved in an O − H⋯O hydrogen bond increases while the blue shift of a C − H bond in a C − H⋯O hydrogen bond decreases when a cooperative effect occurs in ternary complexes as compared to those of the corresponding binary complexes. © 2018 Wiley Periodicals, Inc.  相似文献   

16.
The thermodynamics of guanidinium and boronic acid interactions with carboxylates, alpha-hydroxycarboxylates, and diols were studied by determination of the binding constants of a variety of different guests to four different hosts (7-10). Each host contains a different combination of guanidinium groups and boronic acids. The guests included molecules with carboxylate and/or diol moieties, such as citrate, tartrate, and fructose, among others. The Gibbs free energies of binding were determined by UV/Vis absorption spectroscopy, by use of indicator displacement assays. The receptor based on three guanidinium groups (7) was selective for the tricarboxylate guest. The receptors that incorporated boronic acids (8-10) had higher affinities for guests that included alpha-hydroxycarboxylate and catechol moieties over guests containing only carboxylates or alkanediols. Isothermal titration calorimetry revealed the enthalpic and entropic contributions to the Gibbs free energies of binding. The binding of citrate and tartrate was investigated with hosts 7-10, for which all the binding events were exothermic, with positive entropy. Because of the selectivity of hosts 8-10, a simple boronic acid (14) was also investigated and determined to be selective for alpha-hydroxycarboxylates and catechols over amino acids and alkanediols. Further, the cooperativity of 8 and 9 in binding tartrate was also investigated, revealing little or no cooperativity with 8, but negative cooperativity with 9. A linear entropy/enthalpy compensation relationship for all the hosts 7-10, 14, and the carboxylate-/diol-containing guests was also obtained. This relationship indicates that increasing enthalpy of binding is offset by similar losses in entropy for molecular recognition involving guanidinium and boronic acid groups.  相似文献   

17.
The effect of the ionic environment on the dissociation constant of acid-base indicators is accounted for by using the specific interaction theory (SIT) of Broensted, Scatchard, and Guggenheim. It is shown that the SIT may be applied to evaluate the activity coefficients of organic ions, regardless of their size and structure, if suitable modifications are introduced.  相似文献   

18.
The specific rate constants for the alkaline hydrolysis of formamide were determined in a series of acetone-water mixtures ranging from 0.0% to 70% by weight of the organic solvent and at different temperature, 25°–50°. The rate was found to be larger in water. Addition of acetone appears to cause a change in the solvent structure as concluded from the variation of log k with either 1/D or log[H2O]. The reactivity of the reactants and the activated state seems to be affected differently by a change in the solvent structure. The magnitude of such effect should be considered separately on both the reactants and the activated state in order to account for the increase of the rate as the dielectric constant of the medium is increased.  相似文献   

19.
The idea that hydrogen bond cooperativity is responsible for the structure and reactivity of carbohydrates is examined. Density functional theory and gauge‐including atomic orbital calculations on the known conformers of the α and β anomers of d ‐glucopyranose in the gas phase are used to compute proton NMR chemical shifts and interatomic distances, which are taken as criteria for probing intramolecular interactions. Atom–atom interaction energies are calculated by the interacting quantum atoms approach in the framework of the quantum theory of atoms in molecules. Association of OH1 in the counterclockwise conformers with a strong acceptor, pyridine, is accompanied by cooperative participation from OH2, but there is no significant change in the bonding of the two following 1,2‐diol motifs. The OH6 ... O5 (G?g+/cc/t and G+g?/cc/t conformers) or OH6 ... O4 (Tg+/cc/t conformer) distance is reduced, and the OH6 proton is slightly deshielded. In the latter case, this shortening and the associated increase in the OH6–O4 interaction energy may be interpreted as a small cooperative effect, but intermolecular interaction energies are practically the same for all three conformers. In most of the pyridine complexes, one ortho proton interacts with the endocyclic oxygen O5. Analogous results are obtained when the clockwise conformer, G?g+/cl/g?, detected for the α anomer, and a hypothetical conformer, Tt/cl/g?, are complexed with pyridine through OH6. Generally, the cooperative effect does not go beyond the first two OH groups of a chain. Copyright © 2017 John Wiley & Sons, Ltd.  相似文献   

20.
The self‐assembly of four cyclic D,L‐octapeptides, [‐(D‐Ala‐Gln)4‐], [‐(D‐Val‐Gln)4‐], [‐(D‐Leu‐Gln)4‐], and [‐(D‐Phe‐Gln)4‐], was investigated on the theory level in detail. Based on these cyclic peptides, which contain L‐Gln residues and possess C4 symmetry, a series of oligomers were constructed according to different stacking modes as well as interaction patterns. We employed the semiempirical molecular orbital method AM1 to optimize the structures of all the oligomers, some of which were further studied using density functional method B3PW91/6‐31G to calculate the interaction energies. The studies indicate that when these cyclopeptides aggregate to form oligomers, or even nanotubes, four more hydrogen bonds could form between the sidechains of L‐Gln residues in addition to eight hydrogen bonds formed between the backbones of adjacent two cyclic peptides, a result that would clearly affect the self‐assembling process of cyclic peptides. The main effects can be summarized as follows. First, the dimers of these cyclic peptides with C4 symmetry are more stable than those with D4 symmetry due to their additional H‐bonds between Gln sidechains. Second, for the self‐assembly of the cyclopeptides, there is a competition between parallel and antiparallel stacking modes in lower oligomers such as dimers. However, with an increasing degree of oligomerization, energetically there is an increased possibility for the cyclic peptides to take the parallel stacking mode in assembly. Finally, the synergetic effect of weak interactions is the fundamental driving force for cyclic peptides to form stable nanotubes. © 2005 Wiley Periodicals, Inc. Int J Quantum Chem, 2005  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