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 78 毫秒
1.
The density functional theory and Hartree–Fock methods were used to investigate the proton transfer reaction for a series of model clusters of zeolite/(H2O)n; n=1,2,3, and 4. Without promoted water, the hydrogen-bonded dimer of the water/zeolite system exists as a simple hydrogen-bonded complex, ZOH.(H2O)2, and no proton transfer occurs from zeolite to water. The third promoted water, ZOH(H2O)2H2O, was found to induce a pathway for proton transfer, but at least addition two promoted molecules, ZO(H3O+)H2O(H2O)2, must be involved for complete proton transfer from zeolite to H2O. The results show that the hydronium ion in water cluster adsorbed on zeolite, ZO(H3O+)(H2O)3, can considerably affect the structure and bonding of the hydrogen-bonded dimer of water. The OO distance is contracted from 2.818 Å found in the neutral complex, ZOH(H2O)4, to 2.777 Å for ion-pair complex, ZO(H3O+)(H2O)3. The distance between the oxygen of the hydronium ion and the zeolitic acid site oxygen is predicted to be 2.480 Å which is in good agreement with the experimentally observed value of 2.510 Å. The corresponding density functional adsorption energy of the high coverages of adsorbing molecules on zeolite is calculated to be −9.14 kcal/mol per molecule at B3LYP/6-311+G(d,p) level of theory and compares well with the experimental observation of −8.20 kcal/mol.  相似文献   

2.
Saddle point geometries and barrier heights have been calculated for the H abstraction reaction HO2(2A″)+H(2S) → H2(1Σ+g)+O2(3Σg) and the concerted H approach-O removing reaction HO2 (2A″)+H(2S) → H2O(1A1)+O(3P) by using SDCI wavefunctions with a valence double-zeta plus polarization basis set. The saddle points are found to be of Cs symmetry and the barrier heights are respectively 5.3 and 19.8 kcal by including size consistent correction. Moreoever kinetic parameters have been evaluated within the framework of the TST theory. So activation energies and the rate constants are estimated to be respectively 2.3 kcal and 0.4×109 ℓ mol−1 s−1 for the first reaction, 20.0 kcal and 5.4.10−5 ℓ mol−1 s−1 for the second. Comparison of these results with experimental determinations shows that hydrogen abstraction on HO2 is an efficient mechanism for the formation of H2 + O2, while the concerted mechanism envisaged for the formation of H2O + O is highly unlikely.  相似文献   

3.
We present a study of single color (2 + 1) resonance-enhanced multiphoton ionization (REMPI) of H2O, D2O, and HDO via several Rydberg states lying in the energy range from 80 000 to 86 000 cm−1. Photoelectron spectra (PES) show that the corresponding cations can be vibrationally state-selected for most vibrational states. The exception is the bend of H2O+ and HDO+, where mixing in the REMPI intermediate level results in weak ion intensity and only 50% state purity.  相似文献   

4.
The enthalpy of solvation in sulfolane of the gaseous protonated hydronium dication is estimated as −628 kcal/mol at 298 K. Using literature data, a value ΔHo = −212 kcal/mol is calculated for the reaction H3O+ + H+→ H4O2+ (in sulfolane), supporting the thermodynamic existence of H4O2+ in sulfolane solution, as characterized previously by potentiometric and conductometric titrations. Some aspects of these results are discussed.  相似文献   

5.
A combined MP2 and DFT/B3LYP study of the HXeOH–H2O complex is presented. These computational methods have been used to extract information on the structural, energetical and vibrational properties of the complex. Additionally, we have applied anharmonic vibrational calculations based on the MP2-computed intermolecular potential energy surface. Large perturbations both on the subunit structures and their fundamental vibrational modes are found upon complexation. Large changes of anharmonicity of the HXeOH subunit reflects the perturbation of the molecule's electronic structure. The computed BSSE-corrected interaction energies are −40.23 and −38.94 kJ mol−1 at the CCSD(T)//MP2 and CCSD(T)//B3LYP levels of theory, respectively. The estimated deformation energy contribution to the interaction energy is about 5%, which is very large compared with classical hydrogen-bonded complexes. The topological analysis of the Electron Localization Function (ELF) was applied to study further the hydrogen-bonded interaction between the two complex partners. The obtained interaction pattern suggests that the interaction between HXeOH and H2O is a typical hydrogen bond interaction driven mainly by electrostatic interactions.  相似文献   

6.
Gaussian-2 ab initio calculations were performed to examine the six modes of unimolecular dissociation of cis-CH3CHSH+ (1+), trans-CH3CHSH+ (2+), and CH3SCH2+ (3+): 1+→CH3++trans-HCSH (1); 1+→CH3+trans-HCSH+ (2); 1+→CH4+HCS+ (3); 1+→H2+c-CH2CHS+ (4); 2+→H2+CH3CS+ (5); and 3+→H2+c-CH2CHS+ (6). Reactions (1) and (2) have endothermicities of 584 and 496 kJ mol−1, respectively. Loss of CH4 from 1+ (reaction (3)) proceeds through proton transfer from the S atom to the methyl group, followed by cleavage of the C–C bond. The reaction pathway has an energy barrier of 292 kJ mol−1 and a transition state with a wide spectrum of nonclassical structures. Reaction (4) has a critical energy of 296 kJ mol−1 and it also proceeds through the same proton transfer step as reaction (3), followed by elimination of H2. Formation of CH3CS+ from 2+ (reaction (5)) by loss of H2 proceeds through protonation of the methine (CH) group, followed by dissociation of the H2 moiety. Its energy barrier is 276 kJ mol−1. On both the MP2/6-31G* and QCISD/6-31G* potential-energy surfaces, the H2 1,1-elimination from 3+ (reaction (6)) proceeds via a nonclassical intermediate resembling c-CH3SCH2+ and has a critical energy of 269 kJ mol−1.  相似文献   

7.
Dynamics of Ar+ + H2O collisions, charge transfer (H2O+ formation) and chemical reaction (ArH+ formation), was investigated in crossed-beam experiments in the collision energy range 1–3 eV. The charge transfer occurs both by a simple-electron-jump mechanism and in impulsive intimate collisions. The distributions of relative translational energy of the products show that, most probably, the collisions are only slightly superelastic (by about 0.1 eV), and thus the product H2O+ is highly internally excited. Formation of a small amount of H2O+(B2B2) in very inelastic collisions was observed. The chemical reaction is a direct process with characteristics of the spectator stripping mechanism.  相似文献   

8.
The infrared spectrum of the ionic cluster I(H2O) was recorded from 3170 to 3800 cm−1 by vibrational predissociation spectroscopy. A strong multiplet observed at 3415 cm−1 and a narrow band at 3710 cm−1 were assigned as a hydrogen-bonded OH stretch and free OH stretch respectively, indicating that H2O forms a single hydrogen bond with the iodide anion. Ab initio vibrational frequencies and intensities were computed at the second-order Møller-Plesset (MP2) level for the minimum energy configuration, a nearly linear hydrogen-bonded isomer, and for a low-lying saddlepoint, a symmetric C2v bridged isomer. The spectrum predicted for the hydrogen-bonded isomer agreed well with experiment.  相似文献   

9.
Hydrated strontium borate, SrB4O7·3H2O, has been synthesized and characterized by XRD, FT-IR, DTA-TG and chemical analysis. The molar enthalpy of solution of SrB4O7·3H2O in 1 mol dm−3 HCl(aq) was measured to be (21.15 ± 0.29) kJ mol−1. With incorporation of the previously determined enthalpies of solution of Sr(OH)2·8H2O(s) in [HCl(aq) + H3BO3(aq)] and H3BO3 in HCl(aq), and the enthalpies of formation of H2O(l), Sr(OH)2·8H2O(s) and H3BO3(s), the enthalpy of formation of SrB4O7·3H2O was found to be −(4286.7 ± 3.3) kJ mol−1.  相似文献   

10.
Theoretical investigations on the kinetics of the elementary reaction H2O2+H→H2O+OH were performed using the transition state theory (TST). Ab initio (MP2//CASSCF) and density functional theory (B3LYP) methods were used with large basis set to predict the kinetic parameters; the classical barrier height and the pre-exponential factor. The ZPE and BSSE corrected value of the classical barrier height was predicted to be 4.1 kcal mol−1 for MP2//CASSCF and 4.3 kcal mol−1 for B3LYP calculations. The experimental value fitted from Arrhenius expressions ranges from 3.6 to 3.9 kcal mol−1. Thermal rate constants of the title reaction, based on the ab initio and DFT calculations, was evaluated for temperature ranging from 200 to 2500 K assuming a direct reaction mechanism. The modeled ab initio-TST and DFT–TST rate constants calculated without tunneling were found to be in reasonable agreement with the observed ones indicating that the contribution of the tunneling effect to the reaction was predicted to be unimportant at ambient temperature.  相似文献   

11.
[Re2(Ala)4(H2O)8](ClO4)6 (Re=Eu, Er; Ala=alanine) were synthesized, and the low-temperature heat capacities of the two complexes were measured with a high-precision adiabatic calorimeter over the temperature range from 80 to 370 K. For [Eu2(Ala)4(H2O)8](ClO4)6, two solid–solid phase transitions were found, one in the temperature range from 234.403 to 249.960 K, with peak temperature 243.050 K, the other in the range from 249.960 to 278.881 K, with peak temperature 270.155 K. For [Er2(Ala)4(H2O)8](ClO4)6, one solid–solid phase transition was observed in the range from 270.696 to 282.156 K, with peak temperature 278.970 K. The molar enthalpy increments, ΔHm, and entropy increments,ΔSm, of these phase transitions, were determined to be 455.6 J mol−1, 1.87 J K−1 mol−1 at 243.050 K; 2277 J mol−1, 8.43 J K−1 mol−1 at 270.155 K for [Eu2(Ala)4(H2O)8](ClO4)6; and 4442 J mol−1, 15.92 J K−1 mol−1 at 278.970 K for [Er2(Ala)4(H2O)8](ClO4)6. Thermal decompositions of the two complexes were investigated by use of the thermogravimetric (TG) analysis. A possible mechanism for the thermal decomposition is suggested.  相似文献   

12.
The proton-transfer between ammonia/water and HF/HBr without and with the stimulus of external electric fields(Eext) was investigated with the ab initio calculations. When external electric field is applied, the proton transfer occurs, resulting in ion-paired H4N+X- and H3O+X-(X=Br and F) from hydrogen-bonded complexes in view of the great changes of geometrical structures, dipole moments, frontier molecular orbitals and potential energy surfaces in the critical external electric fields(Ec) of 1.131×107 V/cm for H3N-HBr, 1.378×108 V/cm for H3N-HF, 9.358×107 V/cm for H2O-HBr and 2.304×108 V/cm for H2O-HF, respectively. Furthermore, one or three excess electrons can trigger the proton transfer from H3N-HBr and H3N-HF to H4N+Br- and H4N+F-, while two and four excess electrons can induce the proton transfer from H2O-HBr and H2O-HF to H3O+Br- and H3O+F-, respectively. Compared with that of the analogous NH3/H2O-HCl systems, the strength of Ec of proton transfer increases from HBr to HCl and HF for either H3N-HX or H2O-HX series, which is understandable by the fact that the acidity sequence is HBr>HCl>HF. And the larger of acidity of conjugated acid, the smaller of needed Ec. On the other hand, the Ec for the systems of NH3 with a stronger basicity is generally smaller than that of H2O systems for the same conjugated acid.  相似文献   

13.
The activation barrier for the CH4 + H → CH3 + H2 reaction was evaluated with traditional ab initio and Density Functional Theory (DFT) methods. None of the applied ab initio and DFT methods was able to reproduce the experimental activation barrier of 11.0-12.0 kcal/mol. All ab initio methods (HF, MP2, MP3, MP4, QCISD, QCISD(T), G1, G2, and G2MP2) overestimated the activation energy. The best results were obtained with the G2 and G2MP2 ab initio computational approaches. The zero-point corrected energy was 14.4 kcal mol−1. Some of the exchange DFT methods (HFB) computed energies which were similar to the highly accurate ab initio methods, while the B3LYP hybrid DFT methods underestimated the activation barrier by 3 kcal mol−1. Gradient-corrected DFT methods underestimated the barrier even more. The gradient-corrected DFT method that incorporated the PW91 correlational functional even generated a negative reaction barrier. The suitability of some computational methods for accurately predicting the potential energy surface for this hydrogen radical abstraction reaction was discussed.  相似文献   

14.
The reaction of HOCl + HCl → Cl2 + H2O in the presence of chlorine anion Cl has been studied using ab initio methods. The overall exothermicity is 15.5 kcal mol−1 and this reaction has been shown to have a high activation barrier of 46.5 kcal mol−1. Cl is found to catalyze the reaction via the formation of HOCl·Cl, ClH·HOCl·Cl and Cl·H2) intermediate ion-molecule complexes or by interacting with a concerted four-center transition state of the reaction of HOCl + HCl.  相似文献   

15.
From extraction experiments and γ-activity measurements, the extraction constant corresponding to the equilibrium H3O+(aq) + 1· Na+(nb)  1·H3O+(nb) + Na+(aq) taking place in the two-phase water–nitrobenzene system (1 = hexaethyl p-tert-butylcalix[6]arene hexaacetate; aq = aqueous phase, nb = nitrobenzene phase) was evaluated as log Kex (H3O+, 1·Na+) = −0.6 ± 0.1. Further, the stability constant of the 1·H3O+ complex in water saturated nitrobenzene was calculated for a temperature of 25 °C as log βnb (1·H3O+) = 6.8 ± 0. 2. By using quantum mechanical DFT calculations, the most probable structure of the 1·H3O+ complex species was predicted. In this complex, the hydroxonium ion H3O+ is bound partly to three carbonyl oxygen atoms by strong hydrogen bonds and partly to three alternate phenoxy oxygens by somewhat weaker hydrogen bonds.  相似文献   

16.
The compound [Zn(H2O)4]2[H2As6V15O42(H2O)]·2H2O (1) has been synthesized and characterized by elemental analysis, IR, ESR, magnetic measurement, third-order nonlinear property study and single crystal X-ray diffraction analysis. The compound 1 crystallizes in trigonal space group R3, a=b=12.0601(17) Å, c=33.970(7) Å, γ=120°, V=4278.8(12) Å3, Z=3 and R1(wR2)=0.0512 (0.1171). The crystal structure is constructed from [H2As6V15O42(H2O)]4− anions and [Zn(H2O)4]2+ cations linked through hydrogen bonds into a network. The [H2As6V15O42(H2O)]6− cluster consists of 15 VO5 square pyramids linked by three As2O5 handle-like units.  相似文献   

17.
The geometrical structures of the C3H3 anion are surveyed at the coupled-cluster doubles (CCD) level of theory with the aug-cc-pVDZ basis set. To clarify the CCD geometries, the stable two isomers -- propynl-l-yl 1 and allenyl 2 anions -- are further optimized at the coupled-cluster singles, doubles (triples) (CCSD(T)) level of theory both with the aug-cc-pVDZ and aug-cc-pVTZ basis sets. The final energies are calculated at the CCSD(T) and the complete active space self-consistent field (CASSCF) multi-reference internally contracted CI (MRCI) levels of theory with the aug-cc-pVTZ basis set. At the MRCI level of theory including both the corrections due to the cluster energies (MRCI+Q) and the zero-point vibrational energies, the allenyl anion 2 is about 1.3 kcal mol−1 lower in energy than the propynl-l-yl anion 1. These results contrast with the previous theoretical estimates, where the propynl-l-yl anion 1 is 2-3 kcal mol−1 lower in energy than the allenyl anion 2. The activation energies of the intramolecular hydrogen transfer in the 1 → 2 conversion reactions are 63.5 kcal mol−1 at the MRCI+Q level of theory with the aug-cc-pVTZ basis set including the zero-point energy corrections. The adiabatic electron affinity of the planer propargyl (H2CCCH) radical, which is the global minimum of the C3H3 radical, is calculated to be 0.976 eV (after correction for the zero-point energy changes) at the CCSD(T) level of theory with the aug-cc-pVTZ basis set. The present electron affinity is in fairly good agreement with the experimental one (0.893 eV) observed by Oakes and Ellison.  相似文献   

18.
The molecular structure (equilibrium geometry) and binding energy of the dimethylzinc (DMZn)-hydrogen selenide (H2Se) adduct, (CH3)2Zn:SeH2, have been computed with ab initio molecular orbital and density functional theory (DFT) methods and, where possible, compared with experimental results. The structure of the precursors DMZn and H2Se are perturbed to only a small extent upon adduct formation. (CH3)2Zn:SeH2 was found to be 3 kcal mol−1 less stable than the precursors at the B3LYP/6-311 + G(2d,p)//B3LYP/6-311 + G(2d,p) level of computation, indicating that the (CH3)2Zn:SeH2 adduct is unlikely to be a stable gas-phase species under chemical vapour deposition conditions. Further calculations at the B3LYP/6-311 + G(2d,p)//B3LYP/6-311 + G(2d,p) level of computation suggest that the 1:2 adduct species, (CH3)2Zn:(SeH2)2, is much less stable than the 1:1 adduct and consequently the precursors by 19 kcal mol−1.  相似文献   

19.
Interaction of hydrated proton, H5O2+·(H2O)4, in dichloroethane solutions with diphosphine dioxides (L) having methyl (Ph4Me), ethyl (Ph4Et) and polyoxyethylene chains (Ph4PEG) linking two diphenyl phosphine oxide groups has been investigated. A bulky counter ion: chlorinated cobalt(III) bis(dicarbollide), [Co(C2B9H8Cl3)2], minimizes perturbation of the cation. At low concentrations, Ph4Et and Ph4PEG form anhydrous 1:1 complexes with (P)O–H+–O(P) fragment having very strong symmetrical H-bonds. At these conditions Ph4Me form another compound, H5O2+·L(H2O)2, due to lower PO basicity and optimal geometry of the chelate cycle. At higher concentrations, Ph4Me and Ph4Et form isostructural complexes H5O2+·L2, whereas Ph4PEG forms only a 1:1 complex with proton dihydrate, H3O+·H2O. In excess of free Ph4Me and Ph4Et a water molecule is introduced to the first coordination sphere of H5O2+ and the average molar ratio L/H5O2+ of the complexes exceeds 2. The composition of these complexes as a function of L and its concentration is discussed.  相似文献   

20.
Twenty-two isomers/conformers of C3H6S+√ radical cations have been identified and their heats of formation (ΔHf) at 0 and 298 K have been calculated using the Gaussian-3 (G3) method. Seven of these isomers are known and their ΔHf data are available in the literature for comparison. The least energy isomer is found to be the thioacetone radical cation (4+) with C2v symmetry. In contrast, the least energy C3H6O+√ isomer is the 1-propen-2-ol radical cation. The G3 ΔHf298 of 4+ is calculated to be 859.4 kJ mol−1, ca. 38 kJ mol−1 higher than the literature value, ≤821 kJ mol−1. For allyl mercaptan radical cation (7+), the G3 ΔHf298 is calculated to be 927.8 kJ mol−1, also not in good agreement with the experimental estimate, 956 kJ mol−1. Upon examining the experimental data and carrying out further calculations, it is shown that the G3 ΔHf298 values for 4+ and 7+ should be more reliable than the compiled values. For the five remaining cations with available experimental thermal data, the agreement between the experimental and G3 results ranges from fair to excellent.

Cation CH3CHSCH2+√ (10+) has the least energy among the eleven distonic radical cations identified. Their ΔHf298 values range from 918 to 1151 kJ mol−1. Nevertheless, only one of them, CH2=SCH2CH2+√ (12+), has been observed. Its G3 ΔHf298 value is 980.9 kJ mol−1, in fair agreement with the experimental result, 990 kJ mol−1.

A couple of reactions involving C3H6S+√ isomers CH2=SCH2CH2+√ (12+) and trimethylene sulfide radical cation (13+) have also been studied with the G3 method and the results are consistent with experimental findings.  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