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.
Meta-hybrid density functional theory calculations using M06-2X/6-31+G(d,p) and M06-2X/6-311+G(d,p) levels of theory have been performed to understand the strength of C?Hπ interactions of two possible types for benzene-acetylene, 1,3,5-trifluorobenzene-acetylene and coronene-acetylene complexes. Our study reveals that the C?H...π interaction complex where acetylene located above to the center of benzene ring (classical T-shaped) is the lowest energy structure. This structure is twice more stable than the configuration characterized by H atom of benzene interacting with the π-cloud of acetylene. The binding energy of 2.91 kcal/mol calculated at the M06-2X/6-311+G(d,p) level for the lowest energy configuration (1A) is in very good agreement with the experimental binding energy of 2.7?±?0.2 kcal/mol for benzene-acetylene complex. Interestingly, the C?H...π interaction of acetylene above to the center of the aromatic ring is not the lowest energy configuration for 1,3,5-trifluorobenzene-acetylene and coronene-acetylene complexes. The lowest energy configuration (2A) for the former complex possesses both C?H...π interaction and C?H...F hydrogen bond, while the lowest energy structure for the coronene-acetylene complex involves both π-π and C?H...π interactions. C?H stretching vibrational frequencies and the frequency shifts are reported and analyzed for all of the configurations. We observed red-shift of the vibrational frequency for the stretching mode of the C-H bond that interacts with the π-cloud. Acetylene in the lowest-energy structures of the complexes exhibits significant red-shift of the C?H stretching frequency and change in intensity of the corresponding vibrational frequency, compared to bare acetylene. We have examined the molecular electrostatic potential on the surfaces of benzene, 1,3,5-trifluorobenzene, coronene and acetylene to explain the binding strengths of various complexes studied here.  相似文献   

2.
Increase of the atmospheric concentration of halogenated organic compounds is partially responsible for a change of the global climate. In this work we have investigated the interaction between halogenated ether and water, which is one of the most important constituent of the atmosphere. The structures of the complexes formed by the two most stable conformers of enflurane (a volatile anaesthetic) with one and two water molecules were calculated by means of the counterpoise CP-corrected gradient optimization at the MP2/6–311++G(d,p) level. In these complexes the CH…Ow hydrogen bonds are formed, with the H…Ow distances varying between 2.23 and 2.32 Å. A small contraction of the CH bonds and the blue shifts of the ν(CH) stretching vibrations are predicted. There is also a weak interaction between one of the F atoms and the H atom of water, with the Hw…F distances between 2.41 and 2.87 Å. The CCSD(T)/CBS calculated stabilization energies in these complexes are between ?5.89 and ?4.66 kcal?mol?1, while the enthalpies of formation are between ?4.35 and ?3.22 kcal?mol?1. The Cl halogen bonding between enflurane and water has been found in two complexes. The intermolecular (Cl···O) distance is smaller than the sum of the corresponding van der Waals radii. The CCSD(T)/CBS stabilization energies for these complexes are about ?2 kcal?mol?1.
Figure
Complex between enflurane and water molecules  相似文献   

3.
A survey has been made, by using computer methods, of the types of helices which polypeptide chains can form, taking into account steric requirements and intramolecular hydrogen-bonding interactions. The influence on these two requirements, of small variations in the bond angles of the peptide residues, or of small changes in the overall dimensions of the helix (pitch and residues per turn), have been assessed for the special case of the α-helix. Criteria for the formation of acceptable hydrogen bonds have also been applied to helices of other types, viz., the 3, γ?, ω?, and π-helices. It was shown that the N? H … O and H … O? C angles in hydrogen bonds are sensitive to changes in either the NCαC′ bond angle or in the rotational angles about the N? Cα and Cα? C′ bonds. However, the variants of the α-helix observed experimentally in myoglobin can all be constructed without distortion of the hydrogen bonds. For α-helices, the steric and hydrogen bonding requirements are more easily fulfilled with an NCαC′ bond angle of 111°, rather than 109.5°. The decreased stability observed for the left-handed α-helix relative to the right-handed one for L -amino acids is due essentially only to interactions of the Cβ atom of the side chains with atoms in adjacent peptide units in the backbone, and interactions with atoms in adjacent turns of the helical backbone are not significantly different in the two helices. Restrictions in the freedom of rotation of bulky side chains may have significant kinetic effects during the formation of the α-helix from the “random coil” state.  相似文献   

4.
Self-assembly of melamine-cyanuric acid (MC) leads to urinary tract calculi and renal failure. The hydration effects on molecular geometry, the IR spectra, the frontier molecular orbital, the energy barrier of proton transfer (PT), as well as the stability of MC were explored by density functional theory (DFT) calculations. The intramolecular PT breaks the big π-conjugated ring of melamine or converts the p-π conjugation (:N-C'=O) to π-π conjugation (O=C-N=C') of cyanuric acid. The intermolecular PT varies the coupling between melamine and cyanuric acid from pure hydrogen bonds (Na…HNd and NH…O) to the cooperation of cation…anion electrostatic interaction (NaH+…Nd -) and two NH…O hydrogen bonds. Distinct IR spectra shifts occur for Na…HNd stretching mode upon PT, i.e., blue-shift upon intramolecular PT and red-shift upon intermolecular PT. It is expected that the PT would inhibit the generation of rosette-like structure or one-dimensional tape conformer for the MC complexes. Hydration obviously effects the local geometric structure around the water binding site, as well as the IR spectra of NH…O and N…HN hydrogen bonds. Hydration decreases the intramolecular PT barrier from ~45 kcal mol-1 in anhydrous complex to ~11.5 kcal mol-1 in trihydrated clusters. While, the hydration effects on intermolecular PT barrier is slight. The relative stability of MC varies slightly by hydration due to the strong hydrogen bond interaction between melamine and cyanuric acid fragments.
Graphical Abstract Hydration effect on proton transfer in melamine?cyanuric acid complex
  相似文献   

5.
The interaction of external water molecules with hydrated pyrrole-2-carboxaldehyde PCL/(H2O) n complexes was investigated. The work was supported by both theoretical [DFT/TD-DFT methods using 6-311G++(d,p) basis set in the ground (S0) and excited (S1, S2, S3)states] and experimental [UV-Vis, FTIR and Raman] verification. The focus of the present work was on the weak intermolecular O–H?O, N–H?O–H hydrogen bonded interaction (IerHB) between PCL and external water molecules, and the influence of increasing the number of water molecules to form hydrated PCL/(H2O)n complexes. Effects were observed on different vibrational normal modes and on electronic transition levels. A hydrogen-bonded network of water induces a shift to higher energy in certain normal modes of PCL to form stable PCL/(H2O)n complexes by lowering the barrier energy. Potential energy distribution (PED) analysis indicates a significant charge transfer from PCL to water by creating a water bridge. Hydrogen bonding effects account for the substantial red shift and broadness in νNH, νCO vibrational modes. Water rearrangement turns out to be the main driving force for hydrated complex formation.
Graphical abstract Stability of PCL/(H2O)4 hydarted complex.
  相似文献   

6.
The X-ray structure analysis of a crystalline sample of 2-azabicyclo-[2,2,2]-octanone-3 or 3-isoquinuclidone shows that the molecules of this compound are associated in centrosymmetrical dimers stabilized by two N? H? O?C hydrogen bonds in which the N,H,O atoms are nearly collinear. As a consequence of this interaction, the H atom is shifted from its usual position and the Cα? N? H angle is increased to 125°. Using infrared spectroscopy (νN–H frequency range), it is possible to demonstrate that 3-isoquinuclidone is mainly in a dimeric form when dissolved in an inert solvent such as CCl4 and to observe the dimer-monomer equilibrium on dilution from saturation to a low concentration (0.005 mole/l.). On the contrary, dimers are broken off when operating in a polar medium (acetonitrile, deuterochloroform). In the same experimental conditions, measurements of the J vicinal coupling constant, by nuclear magnetic resonance spectroscopy, afford a concentration-dependent result in the case of CCl4 solutions (increasing from 5.4 to 5.7 Hz when diluting from 0.5 to 0.005 mole/l.) and a constant one (5.8 Hz) in the case of CH3CN or CDCl3 solutions. Then the 0.4-Hz difference can be attributed to geometrical changes in the Hα? Cα? N? H system when dimers are broken off and the valence angle Cα? N? H consequently decreases from 125° to its standard value (about 115°). This experimental observation is consistent with the result of a theoretical analysis performed by the INDO method. Then it seems that the use of the formulas proposed by Karplus to account for the valence angle distorsions in ethane-like systems, in the case of the Hα? Cα? N? H sequence, could yield overstimated corrections.  相似文献   

7.
Infrared spectra of poly-L -alanine in trifluoroacetic acid-chloroform mixtures have been investigated and compared with those of a model amide (N-methylacetamide). The purpose of this work is to determine the nature of peptide-acid specific interactions responsible for the helix-random coil transition of polymer chains. Analysis is made in using amide (A, I, II, III) and acid (νC?O, νOH) vibrations which are specially sensitive to molecular interactions. We examined a model compound to determine the spectral characteristics of the different complexes or species formed between amide and acid. At a low acid concentration, hydrogen-bonded complexes: ? (NH) C?O…?HOOCCF3 (1) are evidenced but no association between amide NH and acid CO groups (complexes A) is observed. For higher acid concentrations complexes (I) are progressively changed into ions pairs and free ions, which result from amide protonation by acid, according to the exothermic equilibrium (I)?? (NH)COH+, ?OOCCF3(II). Amidium and carboxylate bands are localized between 1680–1705 cm?1 and 1620–1625 cm?1, respectively. If the cation band is always clearly seen, the anion band is only observed for the most acidic solutions. For the polymer, a gradual complexation of type (I) is observed for all acid concentrations. From our results, the assumption of an (A) type interaction seems very unlikely but cannot be excluded. Moreover, proton transfer—similar to that observed with a model amide—is never evidenced since, in particular, the amidium band characteristic of protonation is never seen. In contrast to previous investigations, we conclude that the helix-random coil transition of polypeptides is not due to the protonation of the peptide functions. This transition does suggest a strong interaction by hydrogen bonds between polymer and acid molecules.  相似文献   

8.
Two novel ion-pair complexes, [RBzIQl]+[Ni(mnt)2] (mnt2− = maleonitriledithiolate, [RBzIQl]+ = 4-R-benzylisoquinolinium; R = H (1), Cl (2)) have been characterized structurally and magnetically. The anions and [BzIQl]+ cations of 1 form 1D column of alternating between cations and anions via π?π stacking interaction between Ni(mnt)2 plane and isoquinoline ring, and the Ni(mnt)2 anions between adjacent columns exist C?N, C?N, and N?N interaction. The anions and cations of 2 stack into well-segregated columns in the solid state; and the Ni(III) ions form a 1D zigzag chain in a Ni(mnt)2 column through intermolecular Ni?S, S?S, Ni?Ni or π?π interactions. The chain is uniform in 2 with the Ni?Ni distances of 3.784 Å. Magnetic susceptibility measurements for these complexes in the temperature range 1.8-300 K show that 1 exhibits antiferromagnetic coupling behavior, and 2 exhibits unusual magnetic phase transitions around 45 K. The overall magnetic behavior for 2 indicates the presence of antiferromagnetic interaction in the high-temperature phase (HT) and spin gap in the low-temperature phase (LT).  相似文献   

9.
Biosurfactans are amphiphilic compounds synthesized by a wide group of microorganisms and tend to interact with surfaces of different polarities. In the present study we purified and characterized a biosurfactant produced by Dietzia cinnamea KA-1 when cultured by n-hexadecane as sole carbon source. The crude biosurfactant was extracted with ethyl acetate and purified by freezing at–20°C and then silica Gel column chromatography. The purified biosurfactant applied for more characterization using Elemental analysis (CHNS), Fourier Transform Infrared Spectroscopy (FTIR), Mass Spectroscopy (MS) and Nuclear Magnetic Resonance (1H and 13C-NMR) analysis. CHNS analysis showed the presence of C (74.92%) and H (11.63%) but not N or S. Functional groups of OH, CH2, CH3, C=O and aliphatic C?O revealed by FTIR analysis. The presence and position of these groups were confirmed by NMR analysis, and molecular mass of biosurfactant calculated using MS analysis. Finally, the product characterized as a methylated ester compound with molecular formula of C21H42O4. This is the first report of biosurfactant of species D. cinnamea identified as ester, furthermore the ester was found to be in the methylated form.  相似文献   

10.
The decapeptide pBrBz- (Aib)10-OtBu, synthesized by the 5(4H)-oxazolone method, crystallizes in the monoclinic space group C2/c with a = 43.901(2), b = 9.289(2), and c = 34.746(3) A; β = 114.69(3)°; and Z = 8. The crystals contain one molecule of water associated with each peptide. The structure has been solved by the Patterson method and refined to an R value of 0.073 for 6819 observed reflections. The peptide adopts a regular 310-helical structure stabilized by eight N? H …? O?C intramolecular 1 ← 4 (or C10) H bonds. This study has allowed us to characterize this important peptide secondary structure in great detail. The crystal-state conformation agrees well with proposals made on the basis of an ir absorption and 1H-nmr study in solution.  相似文献   

11.
The hydration structure of sodium glycinate (Na+GL?) is probed by the Monte-Carlo multiple minimum (MCMM) method combined with quantum mechanical (QM) calculations at the MP2/6-311++G(d,p) level. In the gas phase, the energy of [Na+GL?]β is more than 30 kJ mol?1 higher than [Na+GL?]α. With higher degrees of hydration, our results indicate that the most stable conformers of [Na+GL?]?(H2O)8 were derived from [Na+GL?]β instead of [Na+GL?]α. The stable conformers determined by the conductor-like polarizable continuum model (CPCM) also show that [Na+GL?]β is more stable than [Na+GL?]α in the liquid phase. By analyzing the hydration process, water…water hydrogen bonding interaction will be more preferable than ion…water interaction as the number of water molecules increases. According to the electronic density at the bond critical point on the Na-X bonds (X?=?O1, O2, N) in the low-energy conformers, Na+GL? will be dissociated as Na+ and GL? in the bulk water, which is not predicted by the CPCM model. The structure features and the charge redistribution of Na+GL? will provide a physical explanation for the weakening Na-O1 interaction.
Graphical Abstract Hydration structure of sodium glycinate from ab initio quantum chemical study
  相似文献   

12.
The molecular structures, relative stability order, and dipole moments of a complete family of 21 planar hypoxanthine (Hyp) prototropic molecular–zwitterionic tautomers including ylidic forms were computationally investigated at the MP2/6–311++G(2df,pd)//B3LYP/6–311++G(d,p) level of theory in vacuum and in three different surrounding environments: continuum with a low dielectric constant (??=?4) corresponding to a hydrophobic interface of protein–nucleic acid interactions, dimethylsulfoxide (DMSO), and water. The keto-N1HN7H tautomer was established to be the global minimum in vacuum and in continuum with ??=?4, while Hyp molecule exists as a mixture of the keto-N1HN9H and keto-N1HN7H tautomers in approximately equal amounts in DMSO and in water at T?=?298.15?K. We found out that neither intramolecular tautomerization by single proton transfer in the Hyp base, nor intermolecular tautomerization by double proton transfer in the most energetically favorable Hyp·Hyp homodimer (symmetry C 2h ), stabilized by two equivalent N1H…O6 H-bonds, induces the formation of the enol tautomer (marked with an asterisk) of Hyp with cis-oriented O6H hydroxyl group relative to neighboring N1C6 bond. We first discovered a new scenario of the keto–enol tautomerization of Hyp?·?Hyp homodimer (C 2h ) via zwitterionic near-orthogonal transition state (TS), stabilized by N1+H…N1? and O6+H…N1? H-bonds, to heterodimer Hyp??·?Hyp (C s ), stabilized by O6H…O6 and N1H…N1 H-bonds. We first showed that Hyp??·?Thy mispair (C s ), stabilized by O6H…O4, N3H…N1, and C2H…O2 H-bonds, mimicking Watson–Crick base pairing, converts to the wobble Hyp?·?Thy base pair (C s ), stabilized by N3H…O6 and N1H…O2 H-bonds, via high- and low-energy TSs and intermediate Hyp?·?Thy?, stabilized by O4H…O6, N1H…N3, and C2H…O2 H-bonds. The most energetically favorable TS is the zwitterionic pair Hyp+?·?Thy? (C s ), stabilized by O6+H…O4?, O6+H…N3?, N1+H…N3?, and N1+H…O2? H-bonds. The authors expressed and substantiated the hypothesis, that the keto tautomer of Hyp is a mutagenic compound, while enol tautomer Hyp? does not possess mutagenic properties. The lifetime of the nonmutagenic tautomer Hyp? exceeds by many orders the time needed to complete a round of DNA replication in the cell. For the first time purine–purine planar H-bonded mispairs containing Hyp in the anti-orientation with respect to the sugar moiety – Hyp?·?Ade syn , Hyp?·?Gua? syn , and Hyp?·?Gua syn , that closely resembles the geometry of the Watson–Crick base pairs, have been suggested as the source of transversions. An influence of the surrounding environment (??=?4) on the stability of studied complexes and corresponding TSs was estimated by means of the conductor-like polarizable continuum model. Electron-topological, structural, vibrational, and energetic characterictics of all conventional and nonconventional H-bonds in the investigated structures are presented. Presented data are key to understanding elementary molecular mechanisms of mutagenic action of Hyp as a product of the adenine deamination in DNA.  相似文献   

13.
The structure and thermodynamic properties of the 2, 4-dinitroimidazole complex with methanol were investigated using the B3LYP and MP2(full) methods with the 6-31++G(2d,p) and 6-311++G(3df,2p) basis sets. Four types of hydrogen bonds [N–H?O, C–H?O, O–H?O (nitro oxygen) and O–H?π] were found. The hydrogen-bonded complex having the highest binding energy had a N–H?O hydrogen bond. Analyses of natural bond orbital (NBO) and atoms-in-molecules (AIM) revealed the nature of the intermolecular hydrogen-binding interaction. The changes in thermodynamic properties from monomers to complexes with temperatures ranging from 200.0 to 800.0 K were investigated using the statistical thermodynamic method. Hydrogen-bonded complexes of 2,4-dinitroimidazole with methanol are fostered by low temperatures.
Figure
Molecular structures and bond critical points of 2,4-dinitroimidazole complexes at MP2(full)/6-311++G(3df,2p) level. Structure and thermodynamic property of the 2,4-dinitroimidazole complex with methanol are investigated using the B3LYP and MP2(full) methods with the 6-31++G(2d,p) and 6-311++G(3df,2p) basis sets. Four types of hydrogen bonds (N–H…O, C–H…O, O–H…O (nitro oxygen) and O–H…π) are found. For the hydrogen-bonded complex having the highest binding energy, there is a N–H…O hydrogen bond. The complex formed by the N–H…O hydrogen bond can be produced spontaneously at room temperature and the equilibrium constant is predicted to be 6.354 and 1.219 at 1 atm with the temperature of 268.0 and 298.15 K, respectively.  相似文献   

14.
An analysis of the geometries of the hydrogen bonds observed by neutron diffraction in thirt-two crystal structures of amino acids shows the following results. Of the 168 hydrogen bonds in the data set, 64 involve the zwitterion groups 
and CO2. Another 18 are from
to sulphate or carbonyl oxygens. The majority, 46, of these
H … O bonds are three-centered (bifurcated). Nine are four-centered (trifurcated). The geometry in which the three-centered hydrogen bond involves both oxygens of the same carboxylate group is not especially favoured. When it does occur, one hydrogen bond is generally shorter and the other longer, than when the bonding involves oxygens on different carboxylate groups. The shortest hydrogen bonds are the OH … O C, from a carboxylic acid hydroxyl to a carboxylate oxygen, and NH … OC when the nitrogen is the ring atom in histidine or proline. Carboxylate groups, on average, accept six hydrogen bonds, with no examples of less than four bonds. The reason for the large number of three-centered
H … OC bonds is therefore a proton deficiency arising from the disparity between the tripled donor property of the
groups and the sextuple, on average, acceptor property of the carboxylate groups. There is good geometrical evidence for the existence of H … O and H … Cl? hydrogen bonds, especially involving the hydrogen atoms on α-atoms.  相似文献   

15.
The molecular and crystal structures of one derivative and three homopeptides (from the di-to the tetrapeptide level) of the chiral, Cα, α-disubstituted glycine Cα-methyl, Cα-benzylglycine [(αMe)Phe], have been determined by x-ray diffraction. The derivative is mClAc-D -(αMe)Phe-OH, and the peptides are pBrBz-[D -(αMe)Phe]2-NHMe, pBrBz-[D -(αMe)Phe]3-OH hemihydrate, and pBrBz-[D -(αMe)Phe]4-OtBu sesquihydrate. All (αMe)Phe residues prefer ?,ψ torsion angles in the helical region of the conformational map. The dipeptide methylamide and the tripeptide carboxylic acid adopt a β-turn conformation with a 1 ← 4 C?O…?H? N intramolecular H bond. The structure of the tripeptide carboxylic acid is further stabilized by a 1 ← 4 C?O…?H? O intramolecular H bond, forming an “oxy-analogue” of a β-turn. The tetrapeptide ester is folded in a regular (incipient) 310-helix. In general, the relationship between (αMe)Phe chirality and helix screw sense is opposite to that exhibited by protein amino acids. A comparison is made with the conclusions extracted from published work on homopeptides from other Cα-methylated α-amino acids. © 1993 John Wiley & Sons, Inc.  相似文献   

16.
The occurrence of type 2 diabetes (T2D) accounts for 90–95 % of all diabetes. Intestine hormone glucagon-like peptide-1 (GLP-1) has an antidiabetic role that enhances insulin secretion and pancreatic β-cell proliferation. GLP-1 is degraded by the enzyme dipeptidyl peptidase-4 (DPP-4) rapidly. Hence, the DPP-4 inhibition has been preferred not only for the treatment but also as a major drug target. Sitagliptin and Diprotin-A are antihyperglycemic agents for the treatment of T2D. However, little is known on the molecular dynamics of DPP-4 and the interaction properties with its ligands, namely Sitagliptin and Diprotin-A. This study has used the latest bioinformatic tools to understand the molecular dynamics and its interaction properties of DPP-4. This study has explored the number of α helices, β strands, β hairpins, Ψ loop, β bulges, β turns, and ? turns and they were 19, 46, 25, 1, 14, 70, and 4, respectively. The highest number of H-bonds was recorded in α helix of domain-1, and the lowest number H-bonds were noted in α helix of domain-2. During interaction between residues, in A- and B-chain, 47 and 48 residues are involved for interaction, and interaction interface area was more in A-Chain (2176 Å2). From DPP-4 and Sitagliptin interaction, three residues in active sites such as Try226, Glu205, and Glu206 were involved in three H-bond formation, while 10 other amino acids (Try547, Try667, Asn710, Val711, His740, Ser630, Ser209, Arg358, Phe357, and Val207) were involved in hydrophobic interactions. In this review, we have shown the importance of bioinformatics as an excellent tool for a rapid method to assess the molecular dynamics and its interaction properties of DPP-4. Our predictions highlighted in this review will help researchers to understand the interaction properties and recognition of interactive sites to design more DPP-4 inhibitors for the treatment of T2D and drug discovery.  相似文献   

17.
The crystal structure of a dipeptide L -leucyl–L -leucine (C12H24N2O3) has been determined. The crystals are monoclinic, space group P21, with a = 5.434(4) Å, b = 15.712(7) Å, c = 11.275(2) Å, β = 100.41(1)°, and Z = 2. The crystals contain one molecule of dimethyl sulfoxide (DMSO) as solvent of crystallization for each dipeptide molecule. The structure has been solved by direct methods and refined to a final R index of 0.059 for 920 reflections (sinθ/λ ? 0.60 Å?1) with I ? 2σ (I). The trans peptide unit shows substantial degree of non-planarity (Δω = 14°). The peptide backbone adopts an extended conformation with torsion angles of ψ1 = 138(1)°, ω1 = 166(1)°, ?2 = ? 149.3(7)°, ψ21 = 164.2(7)°, and ψ22 = ? 15(1)°. For the first leucyl residue, the side-chain conformation is specified by the torsion angles 1χ1 = 176.7(7)°, 1χ21 = 62(1)°, 1χ22 = ? 177.4(8)°; the second leucyl residue adopts a Sterically unfavorable conformation with 2χ1 = 61(1)°, 2χ21 = 97(1)°, and 2χ22 = ?151(1)°. The packing involves head-to-tail interaction of peptide molecules and segregation of polar and nonpolar regions. The DMSO molecule is strongly hydrogen bonded to the terminal NH group. © 1994 John Wiley & Sons, Inc.  相似文献   

18.
The conformation and the interaction of CHF2OCF2CHF2 (desflurane II) with one water molecule is investigated theoretically using the ab initio MP2/aug-cc-pvdz and DFT-based M062X/6-311++G(d,p) methods. The calculations include the optimized geometries, the harmonic frequencies of relevant vibrational modes along with a natural bond orbital (NBO) analysis including the NBO charges, the hybridization of the C atom and the intra- and intermolecular hyperconjugation energies. In the two most stable conformers, the CH bond of the F2HCO- group occupies the gauche position. The hyperconjugation energies are about the same for both conformers and the conformational preference depends on the interaction between the non-bonded F and H atoms. The deprotonation enthalpies of the CH bonds are about the same for both conformers, the proton affinity of the less stable conformer being 3 kcal mol?1 higher. Both conformers of desflurane II interact with water forming cyclic complexes characterized by CH…O and OH…F hydrogen bonds. The binding energies are moderate, ranging from ?2.4 to ?3.2 kcal mol?1 at the MP2 level. The origin of the blue shifts of the ν(CH) vibrations is analyzed. In three of the complexes, the water molecule acts as an electron donor. Interestingly, in these cases a charge transfer is also directed to the non bonded OH group of the water molecule. This effect seems to be a property of polyfluorinated ethers.  相似文献   

19.
Bioethanol is one of the world’s most extensively produced biofuels. However, it is difficult to purify due to the formation of the ethanol–water azeotrope. Knowledge of the azeotrope structure at the molecular level can help to improve existing purification methods. In order to achieve a better understanding of this azeotrope structure, the characterization of (ethanol)5–water heterohexamers was carried out by analyzing the results of electronic structure calculations performed at the B3LYP/6-31+G(d) level. Hexamerization energies were found to range between ?36.8 and ?25.8 kcal/mol. Topological analysis of the electron density confirmed the existence of primary (OH…O) hydrogen bonds (HBs), secondary (CH…O) HBs, and H…H interactions in these clusters. Comparison with three different solvated alcohol systems featuring the same types of atom–atom interactions permitted the following order of stability to be determined: (methanol)5–water > (methanol)6 > (ethanol)5–water > (ethanol)6. These findings, together with accompanying geometric and spectroscopic analyses, show that similar cooperative effects exist among the primary HBs for structures with the same arrangement of primary HBs, regardless of the nature of the molecules involved. This result provides an indication that the molecular ratio can be considered to determine the unusual behavior of the ethanol–water system. The investigation also highlights the presence of several types of weak interaction in addition to primary HBs.
Graphical Abstract Water-ethanol clusters exhibit a variety of interaction types between their atoms, such as primary OH...O (blue), secondary CH...O (green) and H...H (yellow) interactions as revealed by Quantum Chemical Topology
  相似文献   

20.
Crystals of 5‐fluorouridine (5FUrd) have unit cell dimensions a = 7.716(1), b = 5.861(2), c = 13.041(1)Å, α = γ = 90°, β = 96.70° (1), space group P21, Z = 2, ρobs = 1.56 gm/c.c and ρcalc = 1574 gm/c.c The crystal structure was determined with diffractometric data and refined to a final reliability index of 0.042 for the observed 2205 reflections (I ≥ 3σ). The nucleoside has the anti conformation [χ = 53.1(4)°] with the furanose ring in the favorite C2′–endo conformation. The conformation across the sugar exocyclic bond is g+, with values of 49.1(4) and ? 69.3(4)° for Φθc and Φ respectively. The pseudorotational amplitude τm is 34.5 (2) with a phase angle of 171.6(4)°. The crystal structure is stabilized by a network of N–H…O and O–H…O involving the N3 of the uracil base and the sugar O3′ and O2′ as donors and the O2 and O4 of the uracil base and O3′ oxygen as acceptors respectively. Fluorine is neither involved in the hydrogen bonding nor in the stacking interactions. Our studies of several 5‐fluorinated nucleosides show the following preferred conformational features: 1) the most favored anti conformation for the nucleoside [χ varies from ? 20 to + 60°] 2) an inverse correlation between the glycosyl bond distance and the χ angle 3) a wide variation of conformations of the sugar ranging froni C2′–endo through C3′–endo to C4′–exo 4) the preferred g+ across the exocyclic C4′–C5′ bond and 5) no role for the fluorine atom in the hydrogen bonding or base stacking interactions.  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