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 343 毫秒
1.
A theoretical study has been conducted to clarify the discrepancy between ab initio predictions of the entrance channel barrier in the F + H2 reaction and the experimentally observed thresholds for product formation. New results using two different models are in better agreement with the experimental excergicity and predict a zero-point energy corrected barrier of 2.0 kcal mol−1. Estimates of the resulting reaction cross section and thermal activation energy suggest that this value differs from experiment by about 0.5 kcal mol−1. The theoretically predicted isotope effect on the thermal activation energy is larger than either of the recently suggested experimental values, which, however, are subject to large uncertainties.  相似文献   

2.
《Chemical physics letters》1985,122(5):489-492
The rate constant for the bimolecular reaction CO + N2O → CO2 + N2 was determined by comparison of calculated infrared emission profiles of CO2 with those observed in shock-tube experiments in the temperature range 1350–2100 K for CO-N2O-He-Ar mixtures. The rate constant was found to be k1 = 3.2 × 1011exp(−85 kJ/RT) cm3 mol−1 s−1.  相似文献   

3.
Potential-energy surfaces for various channels of the HNO+NO2 reaction have been studied at the G2M(RCC,MP2) level. The calculations show that direct hydrogen abstraction leading to the NO+cis-HONO products should be the most significant reaction mechanism. Based on TST calculations of the rate constant, this channel is predicted to have an activation energy of 6–7 kcal/mol and an A factor of ca. 10−11 cm3 molecule−1 s−1 at ambient temperature. Direct H-abstraction giving NO+trans-HONO has a high barrier on PES and the formation of trans-HONO would rather occur by the addition/1,3-H shift mechanism via the HN(O)NO2 intermediate or by the secondary isomerization of cis-HONO. The formation of NO+HNO2 can take place by direct hydrogen transfer with the barrier of ca. 3 kcal/mol higher than that for the NO+cis-HONO channel. The formation of HNO2 by oxygen abstraction is predicted to be the least significant reaction channel. The rate constant calculated in the temperature range 300–5000 K for the lowest energy path producing NO+cis-HONO gave rise to © 1998 John Wiley & Sons, Inc. Int J Chem Kinet 30: 729–736, 1998  相似文献   

4.
The equilibrium structures, vibrational spectra, and heats of formation for CH3OCl and CH3ClO have been estimated using high levels of ab initio molecular orbital theory. The lowest energy isomer is found to be CH3OCl, and its heat of formation is estimated to be −13.5±2 kcal mol−1, in good agreement with bond additivity estimates. Results for the CH3ClO isomer are presented for the first time, and it is found to be 40.5 kcal mol−1 higher in energy relative to CH3OCl. ©1999 John Wiley & Sons, Inc. Int J Quant Chem 73: 29–35, 1999  相似文献   

5.
The ability of the central M atom of the MCl3 anion, with M=Be, Mg, Ca, Sr, Ba, to engage in a noncovalent bond with an approaching nucleophile is gauged by ab initio methods. The N atom of pyridine forms a M⋅⋅⋅N bond with an interaction energy between 12 and 21 kcal mol−1, even though the π-hole above the M atom is not necessarily positive in sign. Despite a strong Coulombic repulsion between two anions, CN is also able to approach the M atom so as to engage in a metastable complex that is higher in energy than the individual anions. The energy barrier separating this complex from its constituent anion pair is roughly 20 kcal mol−1. Despite the endothermic formation reaction energy of the CN⋅⋅⋅MCl3 complex, the electron topology signals a strong interaction, more so than in pyridine⋅⋅⋅MCl3 with its exothermic binding energy. The dianionic complex is held together largely on the strength of interorbital interactions, thereby overcoming a repulsive electrostatic component. The latter is partially alleviated by the pyramidalization of the MCl3 unit which makes its π-hole more positive. The complex sinks below the separate monomers in energy when the system is immersed in an aqueous medium, with a binding energy that varies from as much as 20 kcal mol−1 for Be down to 1.2 kcal mol−1 for Ba.  相似文献   

6.
《Chemical physics letters》1986,130(3):190-194
Excitation of the C-H overtones of 1,3-butadiene in the presence of HCl to an energy ≈ 16 kcal mol−1 in excess of the activation energy causes the formation of no detectable addition products. This indicates that the rate constant for the formation of 1-chloro-2-butene is at least slower than 2.1× 10−14 cm3 molecule−1 s−1.  相似文献   

7.
The thermodynamic stabilities of P2, P4, and three P8 cage structure were investigated through high‐precision CBS‐Q calculations. The CBS‐Q values for the bond energy of P2 (ΔEo: +115.7 kcal mol−1) and the formation of P4 from P2 (Δ Eo:‐56.6 kcal mol−1) were in excellent agreement with the experimental values (Eo: +117 and ‐56.4 kcal mol−1 respectively). Among the P8 cages, the cubane structure was the least stable (Δ Eo +37 kcal vs. 2×P4). The most stable P8 isomer adopts a cuneane structure resembling S4N4, and is more stable than white phosphorus at T = 0 K (Δ Eo −3.3 kcal mol−1), but still unstable under standard conditions for entropic reasons (Δ Go of +8.1 kcal mol−1 vs. 2×P4). The CBS‐Q energies represent significant revisions (6–20 kcal mol−1) of previous computational predictions obtained by high‐level single method calculations. © 2005 Wiley Periodicals, Inc. Heteroatom Chem 16:453–457, 2005; Published online in Wiley InterScience ( www.interscience.wiley.com ). DOI 10.1002/hc.20119  相似文献   

8.
《Tetrahedron》2019,75(48):130693
The efficiency of bicarbonate molecule (HCO3) as a proton shuttle in the tautomerization and (non)enzymatic CO2 hydration reactions has been investigated with the aid of computational chemistry methods (DFT and ab initio). The results revealed that bicarbonate can decrease the barrier height of tautomerization (keto-enol, azo-hydrazo and imine-amine) more than 70%. This value is around 45% for water molecules. Also, HCO3 can catalyze the CO2 hydration both inside (enzymatic) and outside (nonenzymatic) the active site of human carbonic anhydrases II (HCA II). In the absence of enzyme, bicarbonate molecule can lower the CO2 hydration from ∼50 kcal mol−1 in the gas phase to ∼14 kcal mol−1 in the aqueous media. This reaction maintains its barrier (∼15 kcal mol−1) for bicarbonate-Zn complex in the active site of enzyme; it has been observed that amino acid residues, mainly Thr199 and Glu106, are actively involved in the proton transfer network and facilitate CO2 hydration ability of bicarbonate.  相似文献   

9.
In this work, we examine the strength of various types of individual hydrogen bond (HB) in mixed methanol-water MnWm, (n+m=2 to 7) clusters, with an aim to understand the relative order of their strength, using our recently proposed molecular tailoring-based approach (MTA). Among all the types of HB, it is observed that the OM−H…OW HBs are the strongest (6.9 to 12.4 kcal mol−1). The next ones are OM−H…OM HBs (6.5 to 11.6 kcal mol−1). The OW−H…OW (0.2 to 10.9 kcal mol−1) and OW−H…OM HBs (0.3 to 10.3 kcal mol−1) are the weakest ones. This energetic ordering of HBs is seen to be different from the respective HB energies in the dimer i. e., OM−H…OM (5.0 to 6.0 kcal mol−1)>OW−H…OM (1.5 to 6.0 kcal mol−1)>OM−H…OW (3.8 to 5.6 kcal mol−1)>OW−H…OW (1.2 to 5.0 kcal mol−1). The plausible reason for the difference in the HB energy ordering may be attributed to the increase or decrease in HB strengths due to the formation of cooperative or anti-cooperative HB networks. For instance, the cooperativity contribution towards the different types of HB follows: OM−H…OW (2.4 to 8.6 kcal mol−1)>OM−H…OM (1.3 to 6.3 kcal mol−1)>OW−H…OW (−1.0 to 6.5 kcal mol−1)>OW−H…OM (−1.2 to 5.3 kcal mol−1). This ordering of cooperativity contribution is similar to the HB energy ordering obtained by the MTA-based method. It is emphasized here that, the interplay between the cooperative and anti-cooperative contributions are indispensable for the correct energetic ordering of these HBs.  相似文献   

10.
The kinetics of gas-phase decomposition of methyl isocyanate have been investigated in the range of 427–548°C. Two decomposition routes are followed; the predominant one is a radical-chain process giving CO, H2, and HCN as major products, which has an order of 1.5 and an Arrhenius equation given by log k(L1/2/mol1/2·s) = (13.12 ± 0.06) ? (56,450 ± 1670) cal/mol/2.303 RT. The minor route is the bimolecular formation of N,N′-dimethylcarbodiimide and CO2, which from the low activation parameters Ea = 31.6 kcal, A = 105.30 L1/2/mol1/2·s, and the reaction order of 1.57 appears to be heterogeneous.  相似文献   

11.
Finding novel catalysts for the direct conversion of CO2 to fuels and chemicals is a primary goal in energy and environmental research. In this work, density functional theory (DFT) is used to study possible reaction mechanisms for the conversion of CO2 and C2H6 to propanoic acid over a gold‐exchanged MCM‐22 zeolite catalyst. The reaction begins with the activation of ethane to produce a gold ethyl hydride intermediate. Hydrogen transfers to the framework oxygen leads then to gold ethyl adsorbed on the Brønsted‐acid site. The energy barriers for these steps of ethane activation are 9.3 and 16.3 kcal mol?1, respectively. Two mechanisms of propanoic acid formation are investigated. In the first one, the insertion of CO2 into the Au?H bond of the first intermediate yields gold carboxyl ethyl as subsequent intermediate. This is then converted to propanoic acid by forming the relevant C?C bond. The activation energy of the rate‐determining step of this pathway is 48.2 kcal mol?1. In the second mechanism, CO2 interacts with gold ethyl adsorbed on the Brønsted‐acid site. Propanoic acid is formed via protonation of CO2 by the Brønsted acid and the simultaneous formation of a bond between CO2 and the ethyl group. The activation energy there is 44.2 kcal mol?1, favoring this second pathway at least at low temperatures. Gold‐exchanged MCM‐22 zeolite can therefore, at least in principle, be used as the catalyst for producing propanoic acid from CO2 and ethane.  相似文献   

12.
Amorphous lanthanum carbonate was prepared by hydrolysis of lanthanum isopropoxide using ammonia water in the atmosphere. Lanthanum monoxocarbonate, La2O(CO3)2 · H2O, crystallizes when this amorphous material was washed with hot water. The crystallization and thermal behavior of the crystalline material are studied by X-ray diffraction, thermal analysis, and infrared spectroscopy. The decomposition of La2O(CO3)2 · H2O into type-IA (LaO)2CO3 is observed at 440 to 540°C. Decomposition isotherms are described by the contracting cube equation, the activation energy being 42.6 kcal mol?1. Type-IA (LaO)2CO3 subsequently decomposes to A-type La2O3 at 750 to 870°C. The kinetics is also interpreted in terms of the contracting cube equation, the activation energy being 58.3 kcal mol?1.  相似文献   

13.
The electronic structure and redox properties of the highly oxidizing, isolable RuV?O complex [RuV(N4O)(O)]2+, its oxidation reactions with saturated alkanes (cyclohexane and methane) and inorganic substrates (hydrochloric acid and water), and its intermolecular coupling reaction have been examined by DFT calculations. The oxidation reactions with cyclohexane and methane proceed through hydrogen atom transfer in a transition state with a calculated free energy barrier of 10.8 and 23.8 kcal mol?1, respectively. The overall free energy activation barrier (ΔG=25.5 kcal mol?1) of oxidation of hydrochloric acid can be decomposed into two parts: the formation of [RuIII(N4O)(HOCl)]2+G=15.0 kcal mol?1) and the substitution of HOCl by a water molecule (ΔG=10.5 kcal mol?1). For water oxidation, nucleophilic attack on RuV?O by water, leading to O? O bond formation, has a free energy barrier of 24.0 kcal mol?1, the major component of which comes from the cleavage of the H? OH bond of water. Intermolecular self‐coupling of two molecules of [RuV(N4O)(O)]2+ leads to the [(N4O)RuIV? O2? RuIII(N4O)]4+ complex with a calculated free energy barrier of 12.0 kcal mol?1.  相似文献   

14.
In the title compound [systematic name: 3‐(azaniumylcarbamoyl)pyridinium dichloride], C6H9N3O2+·2Cl, the ions are connected by N—H...Cl hydrogen bonds to form layers and C—H...Cl interactions expand the layers into a three‐dimensional net. The energies of the N—H...Cl interactions range from typical for very weak interactions (0.17 kcal mol−1) to those observed for relatively strong interactions (29.1 kcal mol−1). C—H...Cl interactions can be classified as weak and mildly strong (energies ranging from 2.2 to 8.2 kcal mol−1). Despite the short contacts existing between the parallel aromatic rings of the cations, π–π interactions do not occur.  相似文献   

15.
Using density functional theory methods, we have studied carbon trioxide, its adsorption and dissociation on Ag(100). In the gas phase, two isomers are found, D3h and C2v, with the latter of 2.0 kcal mol?1 lower in energy at the PW91PW91/6?31G(d) level. For CO3 on Ag(100), the calculated adsorption energy is 91.2 and 89.1 kcal mol?1 for the bi‐coord perpendicular and tri‐coord parallel structures, respectively. Upon the adsorption, 0.50 ~ 0.56 electron is transferred from silver to CO3, indicative of significant ionic characters of the adsorbate‐surface bonding. In addition, the geometry of CO3 is largely changed by its strong interaction with silver. For CO3(ad) → O(ad) + CO2(gas), the energy barrier is calculated to be 19.8 kcal mol?1 through the bi‐coord path. The process is endothermic with an enthalpy change of +17.3 ~ +26.7 kcal mol?1 and the weakly chemisorbed CO2 is identified as an intermediate on the potential energy surface. © 2009 Wiley Periodicals, Inc. Int J Quantum Chem, 2010  相似文献   

16.
《Thermochimica Acta》1987,122(2):289-294
The standard enthalpy of formation of potassium metasilicate (K2SiO3), determined by hydrofluoric acid solution calorimetry, was found to be ΔHof,298 = −363.866±0.421 kcal mol−1 (−1522.415±1.762 kj mol−1). The standard enthalpy of formation from the oxides was found to beΔHo298 = −64.786±0.559 kcal mol−1 (−271.065±2.339 kJ mol−1).These experimentally determined data were combined with data from the literature to calculate the Gibbs energies of formation and equilibrium constants of formation over the temperature range of the literature data. The standard enthalpies of formation and Gibbs energies of formation are given as functions of temperature. The standard Gibbs energy of formation is ΔGf,298.150 = −341.705 kcal mol−1 (−1429.694 kJ mol−1).  相似文献   

17.
The hydrogen cyanide exchange mechanism of [Al(HCN)6]3+ has been investigated by DFT calculations (B3LYP/6‐311+G**). The calculations provide theoretical evidence that the hydrogen cyanide exchange proceeds via a limiting dissociative (D) mechanism involving a stable five‐coordinate intermediate [Al(HCN)5 · (HCN)2]3+. The activation energy for the D‐mechanism is 23.4 kcal · mol–1, which is 2.8 kcal · mol–1 lower than for the seven‐coordinate transition state [Al(HCN)7]3+? for the alternative associative (A) pathway. The difference in stability between the two intermediates [Al(HCN)5 · (HCN)2]3+ (12.1 kcal · mol–1) and [Al(HCN)7]3+ (25.7 kcal · mol–1) in comparison to [Al(HCN)6 · (HCN)]3+ is much more pronounced and further supports a limiting dissociative mechanism.  相似文献   

18.
The fluorescence and phosphorescence from benzil in dilute benzene and cyclohexane solutions (2 × 10−4 M) were studied by both conventional luminescence and time-correlated single-photon techniques in the temperature range 8 – 69 °C. The fluorescence (λ = 502 nm) did not show a substantial temperature dependence and was free from thermal and triplet-triplet annihilation delayed contributions at the low concentration used. The phosphorescence (λ = 562 nm) was temperature dependent and its decay was controlled by an activation energy (Ea = 7.4 ± 0.5 kcal mol−1) which was slightly larger than the spectroscopic single-triplet splitting (6.1 kcal mol−1). The photophysical parameters derived from the lifetimes of the two emissions was not consistent with the model of thermal equilibration between S1 and T1.  相似文献   

19.
Triplet‐triplet extinction coefficients for astaxanthin ( I ) and canthaxanthin ( II ) in different deaerated polarity solutions of MeCN and benzene were evaluated by laser flash photolysis at 298 K in the spectral region from 350 to 650 nm by energy transfer method, employing 2‐acetonaphthone as sensitizer. The triplet‐triplet extinction coefficients in MeCN and benzene were different in terms of the carotenoid present. The maximum triplet‐triplet extinction coefficient was 0.1–1.7×105 L·mol−1·cm−1 in different solvents. The rate constants of triplet decay were I : 1.25×1010 L·mol−1·s−1, II : 1.12×1010 L·mol−1·s−1 in MeCN; and I : 1.75×1010 L·mol−1·cm−1, II : 3.27×1010 L·mol−1·s−1 in benzene. The bimolecular rate constants of energy transfer from triplet excited 2‐acetonaphthone to carotenoids were determined from the linear regression of the decay rate constant of 2‐acetonaphthone triplet at varying carotenoid concentrations. The triplet lifetimes of 3AST* and 3CAN* in different solvents were also determined. The results indicated that triplet energy transfer was nearly diffusion‐controlled.  相似文献   

20.
Long B  Long ZW  Wang YB  Tan XF  Han YH  Long CY  Qin SJ  Zhang WJ 《Chemphyschem》2012,13(1):323-329
The formic acid catalyzed gas‐phase reaction between H2O and SO3 and its reverse reaction are respectively investigated by means of quantum chemical calculations at the CCSD(T)//B3LYP/cc‐pv(T+d)z and CCSD(T)//MP2/aug‐cc‐pv(T+d)z levels of theory. Remarkably, the activation energy relative to the reactants for the reaction of H2O with SO3 is lowered through formic acid catalysis from 15.97 kcal mol?1 to ?15.12 and ?14.83 kcal mol?1 for the formed H2O ??? SO3 complex plus HCOOH and the formed H2O ??? HCOOH complex plus SO3, respectively, at the CCSD(T)//MP2/aug‐cc‐pv(T+d)z level. For the reverse reaction, the energy barrier for decomposition of sulfuric acid is reduced to ?3.07 kcal mol?1 from 35.82 kcal mol?1 with the aid of formic acid. The results show that formic acid plays a strong catalytic role in facilitating the formation and decomposition of sulfuric acid. The rate constant of the SO3+H2O reaction with formic acid is 105 times greater than that of the corresponding reaction with water dimer. The calculated rate constant for the HCOOH+H2SO4 reaction is about 10?13 cm3 molecule?1 s?1 in the temperature range 200–280 K. The results of the present investigation show that formic acid plays a crucial role in the cycle between SO3 and H2SO4 in atmospheric chemistry.  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