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 498 毫秒
1.
ABSTRACT

In this study, hydrogen sulfide interaction with pristine, B-, N-, and B&N atom-doped beryllium oxide nanotube (BeONTs) is investigated by the density functional theory (DFT) method. At the first step, we considered different configuration models for the adsorption of H2S on the surface of nanotube and then we selected 12 stable models for this study. The structures of all selected models are optimized and the quantum properties, thermodynamic parameters, natural bond orbitals (NBO), reduced density gradient (RDG), molecular electrostatic potential (MEP), and atom in molecule (AIM) parameters are calculated at the cam–B3LYP level of theory with 6–31G (d) base set. The obtained Eads for the exterior surface of nanotube is exothermic and is in the range –3.15 to –28.34?Kcal mol–1, and that for the interior surface is endothermic and is in the range 19.17 to 27.17?kcal mol–1. The gap energy for pure, B-doped, N-doped, and B&N-doped BeONTs is 10.11, 10.03 (α spin), 10.13 (α spin), and 9.35?eV, respectively. The results of thermodynamic parameters, such as ΔG and ΔH values for the adsorption of H2S, on the surface of B-doped BeONTs are more negative than other models and favorable in thermodynamic approach. The NBO, MEP, NMR, and HOMO–LUMO results confirm that the electron charge transfer occurs from H2S molecule toward BeONTs, as a result the bonding type of H2S?…?BeONTs is weak ionic.  相似文献   

2.
The origin of barriers to internal rotation in ethane-like molecules has been investigated with ab initio molecular orbital theory and natural bond orbital (NBO) analysis. The “hyperconjugative” (vicinal σ → σ*, n → σ*) nature of these interactions, as previously suggested on the basis of semiempirical INDO studies, is confirmed at higher 4–31G, 6–31G* levels of theory. Wave functions lacking such hyperconjugative interactions, but satisfying the Pauli principle, are shown to generally exhibit unphysical barrier behavior (i.e., wrong sign), contrary to the implication of the “Pauli-repulsions” model of Sovers et al. Each trans vicinal σ → σ* interaction is found to contribute about 0.7 kcal/mol to stabilization of the staggered conformer. Geminal interactions and Rydberg orbital contributions are found to be relatively unimportant. The consistency of NBO analysis is exhibited for the methyl series CH3-X (X = CH3, NH2, OH, F) and the family of fluoromethanes, to demonstrate the essential continuity that relates internal rotation barrier forces to static and kinetic anomeric effects, gauche effects, anionic hyperconjugation, bond strengthening in polyhalogenated methanes, (e, 2e) spectroscopic evidence of lone pair delocalization, and other stereoelectronic phenomena.  相似文献   

3.
4.
ABSTRACT

In this study, the stability of the 4-halo-4-H-pyran and 4-halo-4-H-thiopyran molecules in planar and puckered conformations was investigated in the B3LYP/6-311++G(d,p) level of theory. The total energy, dipole moment, the energies of frontier orbitals, and HOMO–LUMO gaps of the molecules were calculated. Natural bond orbital (NBO) analysis was used to illustrate vinylogous anomeric effect on the puckering of ring. The responsible interactions of this effect were determined. The interaction energy, off-diagonal elements and dipole moment values of their NBO were calculated. In addition, total steric exchange energy values of theses interaction were evaluated.  相似文献   

5.
An NMR study of the products of lithiation/methylation of 4,4-dimethyl-4-silathiane S-oxide 1, diastereomers of 2,4,4-trimethyl-4-silathiane S-oxide 2 and 2,4,4,6-tetramethyl-4-silathiane S-oxide 3, as well as 4,4-dimethyl-4-silathiane S,S-dioxide 4 and 2,4,4-trimethyl-4-silathiane S,S-dioxide 5 is reported. The 2-Me group in 2,4,4-trimethyl-4-silathiane S-oxides is always equatorial while the SO group may occupy either the equatorial (major isomer, 2ee) or axial (minor isomer, 2ae) position. 2,4,4,6-Tetramethyl-4-silathiane S-oxide exists in the form of the two isomers, the one with 2-Me, 6-Me and sulfinyl groups all equatorial (3eee), and the other one as an equilibrium mixture of the axial and equatorial sulfoxides with 2-Me axial and 6-Me equatorial (3aae ?s 3eae). The normal Perlin effect (J CHax?<?J CHeq) is found for the CH2 groups in all studied compounds except for the 3- and 5-CH2 groups in 2ae and 5, which show a small reverse Perlin effect (J CHax?>?J CHeq). The experimental findings are interpreted in terms of the σ (C?Hax)?→?σ* (S?O) stereoelectronic effect for the C?H bonds in the 2- and 6-positions, and the buttressing effect of the axial SO group on the C?Hax bonds in the 3- and 5-positions and confirmed by GIAO-B3LYP/6-311G(d,p) theoretical calculations.  相似文献   

6.
Quantum chemical calculations of energies, geometrical structure, intramolecular hydrogen bonding (HB) and vibrational frequencies of 1,1,1-trifluoro-4-mercapto-but-3-ene-2-thione were carried out by the ab initio Hartree–Fock, Moller–Plesset second-order perturbation (MP2) and density functional theory (DFT) methods with 6-311++G** basis set in gas phase and water solution. The nature of the intramolecular hydrogen bond in the most stable chelated conformers has been studied by using the atoms in molecules theory of Bader, which is based on topological properties of the electron density. Natural bond orbital (NBO) analysis was also performed for better understanding of the nature of intramolecular interactions. The influence of the solvent on the stability order of conformers and the strength of intramolecular HB was considered using Tomasi's polarized continuum model. The HOMA, NICS, PDI, ATI, FLU and FLU π indices as well-established aromaticity indicators are examined. The excited-state properties of intramolecular HB in hydrogen-bonded systems have been investigated theoretically using the time-dependent DFT method. The calculated highest occupied molecular orbital (MO) and lowest unoccupied MO with frontier orbital gap are presented. Further verification of the obtained transition state structures was implemented via intrinsic reaction coordinate analysis.  相似文献   

7.
We present laboratory and field measurements of aerosol light extinction ( σep ) using an instrument that employs Cavity Attenuated Phase Shift (CAPS) spectroscopy. The CAPS extinction monitor comprises a light emitting diode (LED), an optical cavity that acts as the sample cell, and a vacuum photodiode for light detection. The particle σep is determined from changes in the phase shift of the distorted waveform of the square-wave modulated LED light that is transmitted through the optical cell. The 3-σ detection limit of the CAPS monitor under dry particle-free air is 3 Mm–1 for 1s integration time. Laboratory measurements of absolute particle extinction cross section ( σext ) using non-absorbing, monodisperse polystyrene latex (PSL) spheres are made with an average precision of ± 3% (2-σ) at both 445 and 632 nm. A comparison with Mie theory scattering calculations indicates that these results are accurate within the 10% uncertainty stated for the particle number density measurements. The CAPS extinction monitor was deployed twice in 2009 to test its robustness and performance outside of the laboratory environment. During these field campaigns, a co-located Multi Angle Absorption Photometer (MAAP) provided particle light absorption coefficient ( σap) at 635 nm: the single scattering albedo ( ω) of the ambient aerosol particles was estimated by combining the CAPS σep measured at 632 nm with the MAAP σap data. Our initial results show the high potential of the CAPS as lightweight, compact instrument to perform precise and accurate σep measurements of atmospheric aerosol particles in both laboratory and field conditions.  相似文献   

8.
β-aminocrotononitrile (1) reacted with either cyanothioacetamide to give (3) or malononitrile to afford an anion (5). Pyridine-2(1H)-thione (4) was obtained by boiling of (3) in ethanol and Et 3N or treatment of (5) with H 2S, respectively. The reaction of anion 5 with isothiocyanates (6) gave N-substituted pyridine-2(1H)-thiones (7). N-Substituted pyridine-2(1H)-thiones (7) can be used for the preparation of pyrido[2,3-d]pyrimidines (8a–e) and (10a–e), or the preparation of pyrido[1,2-a]pyrimidines (12a–d). 1,8-Naphthyridine derivatives (14a–d) and (16a–e) can also be obtained from pyridine-2(1H)-thione (7). Finally, 1,8-naphthyridine derivatives (16a–e) can be used for the preparation of tetracyclic compounds 17a–c and 18a,b.  相似文献   

9.
The effect of the addition of CaX (X=Cl2, F2 and S) on the viscous behavior and structure of CaO–SiO2–Al2O3–MgO–CaX slag was investigated by measuring its viscosity. The viscosity of the slag without CaX gradually decreased with an increase in the C/S ratio because of the depolymerization of the silicate groups in the slag. While the viscosity of the CaX‐bearing slag decreased with an increase in the CaX content, depolymerization was not observed in this case. Three distinct compositional regions for the activation energy of the viscous flow were observed because of the effect of the equilibrium of the polymeric silicate groups. The relaxation effect of the CaX groups on the activation energy was also observed. Raman spectroscopic analysis indicated that the relaxation in the viscosity and activation energy by CaX addition stemmed from the breaking of the NBO‐M2+‐NBO linkage to form NBO‐M2+‐F?, NBO‐M2+‐Cl?, or M2+‐S2?. All these results are discussed in detail with the help of a viscous flow model based on the ionic interactions.  相似文献   

10.
[Ni(fiprdtc)2] (1), [Ni(fiprdtc)(PPh3)(NCS)] (2), [Ni(fiprdtc)(PPh3)2]ClO4 (3), [Zn(fiprdtc)2] (4), [Zn(fiprdtc)2(1,10-phen)] (5) and [Zn(fiprdtc)2(2,2′-bipy)] (6) (f iprdtc=N-furfuryl-N-isopropyldithio- carbamate, 1,10-phen=1,10-phenanthroline and 2,2′-bipy=2,2′-bipyridine) complexes were prepared and characterized by elemental analysis, electronic, IR and NMR spectra and the structure of 2 was determined by single-crystal X-ray crystallography. UV–Vis spectral data of 13 are consistent with the formation of square planar complexes. IR spectra of the complexes show the contribution of the thioureide form to the structure. A single-crystal X-ray structural analysis of 2 proved four-coordinated nickel in a distorted square planar arrangement with a S2PN donor set. Significant asymmetry in Ni–S bond distances was observed in [Ni? S1=2.1655(8); Ni? S2=2.2120(8) Å]. This observation clearly supports the less effective trans of SCN? over PPh3. The observed shielding in N13CS2 chemical shifts of heteroleptic nickel complexes 2 and 3 when compared with homoleptic nickel complex 1 indicates the effect of PPh3 on the mesomeric drift of electron density toward nickel through the thioureide C? N bond. The N13CS2 chemical shift of 5 and 6 are additionally deshielded compared with 4 owing to the increase in coordination number. Complexes were screened for in vitro antibacterial activity and significant activity has been found.  相似文献   

11.
The activity, Φ of AgBr, AgI, PbF2, Ag2S, LiF, and CaF2 in the catalyzed nucleation of poly(ethylene terephthalate) (PET) melts was determined using a nonisothermal differential scanning calorimetry (DSC) technique. A comparison with existing experimental data was made. It is established that the higher the melting temperature of the substrate the lower its activity as a crystallization core in the heterogeneous nucleation of PET. The lateral surface energy, σ, the end surface energy, σe, the adhesion energy, β, and the difference between the surface energies at the substrate/melt, σsf, substrate/deposit, σ*, and the total energy of misfit dislocations, Ed [i.e., σsf - (σ* - Ed)] were calculated. © 1997 John Wiley & Sons, Inc. J Appl Polym Sci 63: 349–353, 1997  相似文献   

12.
The electroreduction of a series of substituted benzoylacetanilides (I) on a dme was investigated in ethanolic Britton—Robinson buffers, and the overall dissociation constants K* of I were evaluated spectrophotometrically under similar conditions. The polarographic curves showed in each case one, two electron cathodic wave, which corresponds to the reduction of the protonated keto form of I. From the pH dependence of the limiting current, the apparent polarographic K′ constants that depend on the protonation of I were determined. Values of the acid dissociation constants of the keto (K1) and enol (K2 tautomers of I were deduced. The values of E12, pK* and pK1 were shown to be linear functions of Hammett substituent constants. The equations of the regression lines obtained are: E12 = 0·284σ—1·379 (r = 0·990) at pH 8·0; pK* = 9·54–2·71σ, (r = 0·998); and pK1 = 9·45–2·946σ, (r = 0·954), respectively. Correlation of pK2 values with substituent constants showed a break at about σ = 0. Controlled potential electrolysis of benzoylacetanilide (Id) produces 3-phenyl-3-hydroxypropionanilide (IIId). A mechanism for the polarographic reduction of I is proposed.  相似文献   

13.
The series of bimetallic complexes, [(η5‐C5Me5)Zr(Me)2]2 [N(t‐Bu)C(Me)N (CH2)n NC(Me)N(t‐Bu)] 3 (n=8), 4 (n=6), and 5 (n=4) were prepared in high yield through a simple, one‐pot synthesis involving 2 equiv. of in situ generated (η5‐C5Me5)Zr(Me)3 and the corresponding bis‐carbodiimide, (t‐Bu)NCN (CH2)n NCN(t‐Bu). Compounds 3 – 5 were found to be highly isoselective for the living Ziegler–Natta polymerization of propene upon 100% activation using 2 equiv. of the borate co‐initiator, [PhNHMe2] [B(C6F5)4] ( 2 ), with the degree of stereoselectivity decreasing slightly as the two metal centers are brought closer together [cf., 3 (σ=0.92)> 4 (σ=0.91)> 5 (σ=0.89)]. Under conditions of sub‐stoichiometric activation by 2 , all three bimetallic initiators, 3 – 5 , were found to engage in degenerative transfer living Ziegler–Natta polymerization involving rapid and reversible methyl group transfer between active, (cationic) and dormant, (neutral) methyl, polymeryl zirconium centers. Under these conditions, the frequency of mr triad stereoerror incorporation into the polypropene (PP) microstructure decreases as the two metal centers are brought closer together as a result of increasing barriers for metal‐centered epimerization within the neutral metal site due to correspondingly greater non‐bonded steric interactions vis‐à‐vis mononuclear 1 .  相似文献   

14.
Transition metal complexes of Mn(II), Fe(III), Cr(III) and Zn(II) metal ions with a general formulas [Mn(L)2(Cl)2]·4H2O (I), [Fe(L)2(Cl)2]·Cl·6H2O (II), [Cr(L)2(Cl)2]·Cl·6H2O (III) and [Zn(L)2(Cl)2]·2H2O (IV) where L = 4-acetylpyridine thiosemicarbazone, have been synthesized and interpreted using CHN elemental analysis, magnetic susceptibility measurements, molar conductance, thermal analysis and spectroscopic techniques; i.e., infrared, electronic UV/vis, 1H-NMR and mass. The manganese(II), ferric(III), chromium(III) and zinc(II) complexes have octahedral geometry. The molar conductance measurements reveal that the Mn(II) and Zn(II) chelates are non-electrolytes but Fe(III) and Cr(III) have an electrolytic behavior. The IR spectra show that the 4-acetylpyridine thiosemicarbazone free ligand is coordinated to the metal(II) chlorides as a neutral bidentate ligand through both of the lone pair of electrons of the C=N azomethine group and C=S group. X-ray powder diffraction gives an impression that the resulting complexes are amorphous and different from the start materials. The thermogravimetric studies indicate that uncoordinated water molecules are lost in the first and second decomposition steps. The activation thermodynamic parameters E*, ΔH*, ΔS* and ΔG* are estimated from the differential thermogravimetric analysis (DTG) curves using Horowitz–Metzger (HM) and Coats–Redfern (CR) methods. The ligand and its complexes have been screened for antibacterial and antifungal activities against two bacteria; i.e., Escherichia coli (Gram −ve) and Bacillus subtilis (Gram +ve) and two fungi, i.e., tricoderma and penicillium activities).  相似文献   

15.
The adsorption of water on a C2S surface initiates belite to hydrate. In the present work, the adsorption behavior of single water molecule on a β‐C2S (100) surface is explored using density functional theory (DFT) due to the lack of alternative approaches for direct observation. Four possible calcium atom sites on the β‐C2S (100) surface slab are considered in our calculations. The results show that water can adsorb on the 2 five‐coordinated calcium sites only via molecular adsorption with adsorption energies of 0.59 and 0.85 eV, respectively, but can dissociate on the other 2 six‐coordinated calcium sites with higher adsorption energies of 0.96 and 0.99 eV, respectively. The energy barriers to the dissociative adsorption of water at the Ca(III) site(0.10 eV) is much lower than that at the Ca(IV) site, indicating that water prefers to adsorb and dissociate on Ca(III) sites. The dissociative adsorption of water causes more obvious surface calcium shifts and Si–O bond length increases than molecular adsorption. The dissociative adsorption of a water molecule changes the electron distribution, and the overlap between Ca 2p and O 2s orbitals leads to new Ca–O bond formations.  相似文献   

16.
《Ceramics International》2022,48(2):1765-1770
Perovskite lattice was tailored by introducing site vacancies and mixed anion composition, to produce Sr0.83Li0.17Ta0.83O1.88N0.74 (Li02N). Further, Li02N was converted to a defect oxide Sr0.83Li0.17Ta0.83O3 (Li02O) by applying an optimized treatment: heating in air at 1173 K for 2 h. According to the neutron Rietveld refinement, Li02N and Li02O are tetragonal and orthorhombic, respectively, where the lattice volume of Li02O is significantly smaller than that of Li02N. The ionic conductivity (σion) of Li02N and Li02O was evaluated by the ac impedance spectroscopy and the equivalent circuit analysis. Both Li02N (σion = 10?5.5 S/cm at 671 K) and Li02O (σion = 10?6.2 S/cm at 667 K) exhibited an Arrhenius behavior of ionic conductivity with activation energies of 0.87 eV and 0.75 eV, respectively. It is interpreted that the nitride component enhances the ionic conduction of Li02N, while the vacancy of the anion lattice makes an opposite effect.  相似文献   

17.
The electrochemical behaviour of the series of ten [Rh(RCOCHCOR′)(P(OPh)3)2] complexes with R, R′ = CF3, CF3 (1), CF3, CH3 (2), CF3, Ph (C6H5) (3), CF3, Fc (ferrocenyl = (C5H5)Fe(C5H4)) (4), CH3, Ph (5), CH3, CH3 (6), Ph, Ph (7), Fc, CH3 (8), Fc, Ph (9) and Fc, Fc (10) were studied in acetonitrile containing 0.100 mol dm−3 tetra-n-butylammonium hexafluorophosphate as supporting electrolyte utilizing a glassy carbon working electrode. Results are consistent with Rh(I) being first oxidized in an electrochemically irreversible two-electron transfer process at peak anodic potentials ranging Epa(Rh) = 0.124–0.881 V vs. Fc/Fc+. For the ferrocene-containing complexes (4), and (8)–(10) the rhodium oxidation was followed by the electrochemically reversible oxidation of the ferrocenyl group in a one-electron transfer process at a slightly more positive potential. Relationships were established between the electrochemical quantity Epa(Rh) and kinetic parameter log k2 as well the sum of experimental group electronegativities (Gordy Scale) of the R and R′ groups (χR + χR′), the Hammett σ values (σR + σR′) and the Lever ligand parameter EL for the [Rh(RCOCHCOR′)(P(OPh)3)2] complexes: Epa(Rh) (vs. Fc/Fc+/V) = 0.31 (χR + χR′)–1.09 = 0.56 (σR + σR′) + 0.28 = SMEL) + (IM − 0.66 V) = −0.23 log k2 − 0.03 (k2 = second order rate constant for the oxidative addition of methyl iodide to rhodium). A profound shift of Epa(Rh) to a more positive potential was observed for Rh(I) substrates containing β-diketonato ligands with increasing electronegative substituents R and R′. An exponential dependence of Epa(Rh) on the pKa of the β-diketone was obtained.  相似文献   

18.
Zhang  Nan  Qian  Guoyu  Wang  Zhi  Wei  Kuixian  Ma  Wenhui  Gong  Wei 《SILICON》2020,12(5):1145-1156

In present work, the role of the oxygen potential (PO2) and oxygen ion (O2−) concentration for removing phosphorus (P) during CaO–SiO2–Al2O3–FexO slag refining was studied by on-line measurement of oxygen activity in molten silicon (Si), FactSage calculation, Raman spectroscopy and nuclear magnetic resonance (NMR) spectroscopy. The results show that the addition of FeO from 0 to 9.25 wt% in slag can increase the activity of dissolved oxygen (a[O]) in Si and the mole fraction of O2− in slag. Moreover, the increase of O2− concentration leads to the increase of non-bridge oxygen (NBO). The value of LP (the partition ratio of phosphorous between slag and Si shows a first increase and then decrease trend and reaches a maximum value of 1.95 at 5 ± 0.1 wt% FeO. It is believed that the increase of a[O] and NBO can promote the removal of P as FeO content is less than 5 ± 0.1 wt%. the chain structure unit (Q2) of silicate network as the main intermediate structure to capture PO43− from the charge compensation of P2O5 by O2− to form the sheet structure unit Q3(Si and P). When FeO content is increased to more than 5 ± 0.1 wt%, LP value gradually decreases although the values of NBO and a[O] are increasing. NBO plays a leading role in this process, it can be speculated that more NBO can depolymerize the Q3 (Si and P) to destroy the stability of P in silicate network. As a result, a mount of PO43− is present at the interface to prevent the oxidation of phosphorous, which leads to the decrease of LP value.

  相似文献   

19.
The geometries, interaction energies, and bonding properties of cationic chalcogen bonds are studied in binary complexes XF2Y+?NCZ (X═H, CN, F; Y═S, Se; Z═H, Cl, Br). The nature of these interactions is studied by a vast number of methods, including molecular electrostatic potential (MEP), Noncovalent Interaction Index (NCI), quantum theory of atoms in molecules (QTAIM), and natural bond orbital (NBO) analyses. The interaction energies of these complexes vary between ?20.94?kcal/mol in HF2S+?NCH and ?33.72?kcal/mol in F3Se+?NCBr. According to the QTAIM analysis, all these cationic chalcogen bonds are classified as a closed-shell interaction with a partial covalent character. Moreover, cooperative effects between cationic chalcogen bond and hydrogen or halogen bond interactions are studied in ternary XF2Y+?NCZ?NH3 complexes. These cooperative effects are analyzed in terms of the parameters derived from the QTAIM and NBO analyses, and electron density difference plots.  相似文献   

20.
Summary The reaction of HOC6H4CH2CN , N3P3(O2C12H8)2(OC6H4CH2CN)2 and – [{NP(O2C12H8)}0.8 {NP (OC6H4CH2CN)2}0.18 ]n with Cp*Fe(dppe)I in dichloromethane solution and in the presence of TIPF6 affords the new compounds [Cp*Fe(dppe)NCCH2C6H4OH][PF6]1 Cp* = C5(CH3)5 [{N3P3 (O2C12H8)2(OC6H4CH2CN•Cp*Fe(dppe))2] [PF6]2 2 and the copolymer [{NP{(O2C12H8}0.8 {NP(OC6H4CH2CN•[Cp*Fe(dppe)][PF6])2}O.8]n 3 respectively. The σ coordination of the Cp*Fe(dppe) fragment toward the nitrile group is indicated by spectroscopic data. The copolymer 3 is soluble in several organic solvents with no significant cross-linking and has a Mw on the order of 1.260.000. Thermal effects of the incorporation of the organometallic fragment to the copolymer were investigated using differential scanning calorimetry (DSC) and therrnogravimetric analysis (TGA). Received: 7 May 2002/Revised version: 21 March 2003/ Accepted: 27 March 2003 Correspondence to C. Diaz  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