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 750 毫秒
1.
Systematic calculations of the structures of the H2CO, F2CO, Cl2CO, HClCO, HFCO, and FClCO molecules in the S0 and T1 states were performed using the B3LYP and MP2 methods with different AO basis sets and also at the CCSD(T)/cc-pV(T+d)Z level of theory. The saturation of the correlation consistent sequence of basis sets cc-pV(N+d)Z (N = D, T, Q, and 5) and aug-cc-pV(N+d)Z (N = D, T, and Q) was studied. Recommendations for choosing the calculation method are given. The relativistic corrections were estimated. The influence of the number and type of halogen atoms on the geometric parameters of the molecules in the S0 and T1 states and the heights of inversion barriers in the T1 state was investigated. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 12, pp. 2625–2635, December, 2005.  相似文献   

2.
Ab initio calculations have been performed on the complexes of CF2Cl2 with NO and SO2, and a set of stable configurations for CF2Cl2–NO and CF2Cl2–SO2 were found with no imaginary frequencies by the MP2 method. In addition, the binding energy and the NBO analysis were used to evaluate the relative stability of the complexes. The calculated results indicate that the weak interactions in the CF2Cl2–NO and CF2Cl2–SO2 systems involved are enhanced with the increase of the number of non-covalent bonds. Further studies predict that the CF2Cl2–SO2 system may play a more important role than the CF2Cl2–NO system in environmental problem because the former offers a stronger interaction than the latter. Furthermore, the non-covalent binding interactions of Cl···N and Cl···O for CF2Cl2–NO system, Cl···O, Cl···S and F···S for CF2Cl2–SO2 system, are the dominant forces, which seem to be very significant as a driving force influencing the arrangement of molecules, especially in CF2Cl2–SO2 system.  相似文献   

3.
Ab initio and Rice–Ramsperger–Kassel–Marcus theories are carried out to study the potential energy surface and the energy‐dependent rate constants and branching ratios of the products for O(1D) + CH3CHF2 reaction. Optimized geometries and vibrational frequencies have been obtained by MP2/6‐311G(d,p) method. The main products of the title reaction are CH3CFO + HF, CH2CFOH + HF, and CH3 + CF2OH at lower collision energy; and CH3 + CF2OH, CH3CF2 + OH are the main products at higher collision energy. CHF2 + CH2OH are the main products in the whole range of collision energy. © 2011 Wiley Periodicals, Inc. Int J Quantum Chem, 2012  相似文献   

4.
The CW–CO2 laser induced reaction of CF2HCH3 (1) with Cl2 yields vinylidene fluoride (2) as the main product; other products are CH3CF2Cl (3), CF2ClCH2Cl (4), CF2 = CHCl (5), CF2 = CCl2 (6) and CF2 Cl2 (7). The yield dependence of 2 on the CF2HCH3/Cl2 ratio, the laser irradiation time, the laser power and the pressure of the gaseous reactants have been investigated. Furthermore, the TEA–CO2 laser induced reaction of 1 with Cl2. the CW–CO2 laser induced reaction of 2 and 2 with Cl2 have also been studied in order to gain more mechanistic insights for this complicated reaction system. Apparently, CF2ClCH3 but not CF2HCH2Cl. is the main precursor to 2. Interestingly, it has been found that the relatively strong double bond of CF2 = CH2 can be broken by laser irradiation. The possibility of applying this laser methodology to the production of vinylidene fluoride has been discussed.  相似文献   

5.
The CO_2 laser induced room temperature reactions of CH_3CF_2H or another protium-donorCH_3CHClCH_3 with chlorine-atom donors (Z--Cl) CFCl_2CF_2Cl, CF_3CCl_3, CFCl_3 or CF_2Cl_2, havebeen investigated. Some of these reactions can yield two important monomers (CF_2=CH_2 andCF_2=CFCl) for fluoropolymers simultaneously. The yield dependence of these two alkenes on experi-mental conditions has been studied. A laser-initiated chain process is supported by identifica-tion of Z--H intermediates in these reactions.  相似文献   

6.
The reaction between Cl2Te(NSO)2, Cl6Te2N2S and Cl2Te(N=S=N)2TeCl2 with MCl3 provided the compounds [(Cl2Te)2N+][MCl4] (M = Ga, Al, Fe). Treating Cl6Te2N2S with M′Cl3 yielded besides [(Cl2Te)2N+][M′Cl4] (M′ = Al, Fe) the sulfur containing compound [ClTeNSNS+][M′Cl4]. The structure for [ClTeNSNS+][FeCl4] was established by an X‐ray structure analysis. With Te(NSO)2 and CF3SCl, via Cl2Te(NSO)2, the known compound Te2NCl5 was formed. Tetrafluoroditelluradiazetidine was obtained from TeF4 and [(CH3)3Si]2NH which on treating with (CH3)3SiCl provided the corresponding chloroderivative. In addition metathetical reaction between Cl2TeNSNS and CF3C(O)OAg yielded [CF3C(O)O]2TeSNSN. Similarly (CH3)2Te(NSO)2–xClx (x = 0,1) and (CH3)2Te(NCO)2 were made from (CH3)2TeCl2 and AgNSO or AgNCO, respectively. Halogination of Cl2Te(N=S=N)2TeCl2 with Cl2 or Br2 yielded Cl6Te2N2S and Cl4Br2Te2N2S. The bromoderivate was also prepared from Cl2Te(NSO)2 and Br2. AgNSO was synthesized by treating CF3C(O)OAg with (CH3)3SiNSO. Two other synthons (CF3Se)2Te and (CF3S)2Se were obtained from CF3SeCl and Na2Te and from Hg(SCF3)2 plus SeCl4, respectively.  相似文献   

7.
In this article, we report our detailed mechanistic study on the reactions of cyclic-N3 with NO, NO2 at the G3B3//B3LYP/6-311+G(d) and CCSD(T)/aug-cc-pVTZ//QCISD/6-311+G(d)+ZPVE levels; the reactions of cyclic-N3 with Cl2 was studied at the G3B3//B3LYP/6-311+G(d) and CCSD(T)/aug-cc-pVTZ//QCISD/6-31+G(d)+ZPVE levels. Both of the singlet and triplet potential-energy surfaces (PESs) of cyclic-N3 + NO, cyclic-N3 + NO2 and the PES of cyclic-N3 + Cl2 have been depicted. The results indicate that on singlet PESs cyclic-N3 can undergo the barrierless addition–elimination mechanism with NO and NO2 forming the respective dominant products N2 + 1cyclic-NON and 1NNO(O) + N2. Yet the two reactions on triplet PESs are much less likely to take place under room temperature due to the high barriers. For the cyclic-N3 + Cl2 reaction, a Cl-abstraction mechanism was revealed that results in the product cyclic-N3Cl + Cl with an overall barrier as high as 14.7 kcal/mol at CCSD(T)/aug-cc-pVTZ//QCISD/6-31+G(d)+ZPVE level. So the cyclic-N3 radical could be stable against Cl2 at low temperatures in gas phase. The present results can be useful for future experimental investigation on the title reactions.  相似文献   

8.
We have calculated the thermochemical parameters for the reactions H(2)SO(4) + H(2)O <--> H(2)SO(4).H(2)O and H(2)SO(4) + NH(3) <--> H(2)SO(4).NH(3) using the B3LYP and PW91 functionals, MP2 perturbation theory and four different basis sets. Different methods and basis sets yield very different results with respect to, for example, the reaction free energies. A large part, but not all, of these differences are caused by basis set superposition error (BSSE), which is on the order of 1-3 kcal mol(-1) for most method/basis set combinations used in previous studies. Complete basis set extrapolation (CBS) calculations using the cc-pV(X+d)Z and aug-cc-pV(X+d)Z basis sets (with X = D, T, Q) at the B3LYP level indicate that if BSSE errors of less than 0.2 kcal mol(-1) are desired in uncorrected calculations, basis sets of at least aug-cc-pV(T+d)Z quality should be used. The use of additional augmented basis functions is also shown to be important, as the BSSE error is significant for the nonaugmented basis sets even at the quadruple-zeta level. The effect of anharmonic corrections to the zero-point energies and thermal contributions to the free energy are shown to be around 0.4 kcal mol(-1) for the H(2)SO(4).H(2)O cluster at 298 K. Single-point CCSD(T) calculations for the H(2)SO(4).H(2)O cluster also indicate that B3LYP and MP2 calculations reproduce the CCSD(T) energies well, whereas the PW91 results are significantly overbinding. However, basis-set limit extrapolations at the CCSD(T) level indicate that the B3LYP binding energies are too low by ca. 1-2 kcal/mol. This probably explains the difference of about 2 kcal mol(-1) for the free energy of the H(2)SO(4) + H(2)O <--> H(2)SO(4).H(2)O reaction between the counterpoise-corrected B3LYP calculations with large basis sets and the diffusion-based experimental values of S. M. Ball, D. R. Hanson, F. L Eisele and P. H. McMurry (J. Phys. Chem. A. 2000, 104, 1715). Topological analysis of the electronic charge density based on the quantum theory of atoms in molecules (QTAIM) shows that different method/basis set combinations lead to qualitatively different bonding patterns for the H(2)SO(4).NH(3) cluster. Using QTAIM analysis, we have also defined a proton transfer degree parameter which may be useful in further studies.  相似文献   

9.
The accurate estimation of S-O bond dissociation enthalpies (BDE) of sulfoxides by computational chemistry methods has been a significant challenge. One of the primary causes for this challenge is the well-established requirement of including high-exponent d functions in the sulfur basis set for accurate energies. Unfortunately, even when high-exponent d functions were included in Pople-style basis sets, the relative strength of experimentally determined S-O BDE was incorrectly predicted. The aug-cc-pV(n+d)Z basis sets developed by Dunning include an additional high-exponent d function on sulfur. Thus, it was expected that the aug-cc-pV(n+d)Z basis sets would improve the prediction of sulfoxide S-O BDE. This study presents the S-O BDE predicted by B3LYP, CCSD, CCSD(T), M05-2X, M06-2X, and MP2 combined with aug-cc-pV(n+d)Z, aug-cc-pVnZ, and Pople-style basis sets. The accuracy of these predictions was determined by comparing the computationally predicted values to the experimentally determined S-O BDE. Values within experimental error were obtained for dialkyl sulfoxides when the S-O BDEs were estimated using an isodesmic oxygen transfer reaction at the M06-2X/aug-cc-pV(T+d)Z level of theory. However, the S-O BDE of divinyl sulfoxide was overestimated by this method.  相似文献   

10.
A laser flash photolysis–resonance fluorescence technique has been employed to investigate the kinetics and mechanism of the reaction of electronically excited oxygen atoms, O(1D), with CF2HBr. Absolute rate coefficients (k1) for the deactivation of O(1D) by CF2HBr have been measured as a function of temperature over the range 211–425 K. The results are well described by the Arrhenius expression k1(T) = 1.72 × 10?10 exp(+72/T) cm3molecule?1 s?1; the accuracy of each reported rate coefficient is estimated to be ±15% (2σ). The branching ratio for nonreactive quenching of O(1D) to the ground state, O(3P), is found to be 0.39 ± 0.06 independent of temperature, while the branching ratio for production of hydrogen atoms at 298 K is found to be 0.02?0.02+0.01. The above results are considered in conjunction with other published information to examine reactivity trends in O(1D) + CF2XY reactions (X,Y = H, F, Cl, Br). © 2001 John Wiley & Sons, Inc. Int J Chem Kinet 33: 262–270, 2001  相似文献   

11.
The computational investigations are carried out on heterodimers containing sulfur tetroxide (SO4(C2V)) with the nitrous oxide (NNO) through MP2/cc–pVDZ and MP2/aug–cc–pVTZ//MP2/cc–pVDZ levels. Eight heterodimers are located on the potential energy surface of SO4(C2V)–NNO system. Binding energies of heterodimers in the SO4(C2V)–NNO system corrected with BSSE and ZPE are in the range of 1.17–7.90 kJ/mol. The calculated results reveal that the individual interaction of NNO terminal nitrogen atom with one of oxygen atoms of OSO ring in the SO4(C2V) monomer leads to the formation of the more stable heterodimer of SO4(C2V)–NNO system. The atoms in molecules theory were applied to analyze the nature of intermolecular interactions.  相似文献   

12.
CF3O2CF3 was photolyzed at 254 nm in the presence of CO in 760 torr N2 or air at 296 K in a static reactor. In N2, the products CF3OC(O)C(O)OCF3 and CF3OC(O)O2C(O)OCF3 were detected by FTIR spectroscopy. In air, the only observed products were CF2O and CO2 and a chain process, initiated by CF3O, was invoked for the conversion of CO to CO2. From both product studies, a mechanism for the CF3O initiated oxidation of CO was derived, involving the addition reaction CF3O2 + CO → CF3OC(O). The rate constant for the reaction CF3O + CO at 296 K at a total pressure of 760 torr air was determined to be k(CF3O + CO) = (5.0 ± 0.9) × 10−14 cm3 molecule−1 s−1. © 1997 John Wiley & Sons, Inc.  相似文献   

13.
The reliability of the two-layer own N-layered integrated molecular orbital and molecular mechanics (ONIOM) method was examined for the SN2 reaction CH(4–n)Cln+OH. In the ONIOM calculation, only the methyl chloride and OHwere treated at a high level and the effect of polychlorination was taken into account only at a low level. The ONIOM results were compared with the target CCSD(T)/aug-cc-pVDZ//MP2/aug-cc-pVDZ results obtained by Borisov etal. [(2001) J. Phys. Chem. A 105:7724]. The ONIOM[MP2/aug-cc-pVDZ:B3LYP/6-31+G(d)] was found to reproduce well the target geometry and energy at the MP2/aug-cc-pVDZ level. The single-point improved energetics at the ONIOM[CCSD(T)/aug-cc-pVDZ:MP2/6-31+G(d)] is found to give results nearly as accurate as the target CCSD(T)/aug-cc-pVDZ//MP2/aug-cc-pVDZ results. The substantially reduced cost, 20% for optimization and 5% for single-point improved energy of the target cost for n=4, as well as small errors suggest that ONIOM is a powerful tool for accurate potential-energy surfaces of the reaction of large polyhalohydrocarbons.  相似文献   

14.
The tris‐bidentate complex [Ni(C12H8N2)2(CO3)] was synthesized by the reaction of Na2CO3, 1,10‐phenanthroline (phen = C12H8N2) and NiSO4 · 6 H2O in the presence of succinic acid in a CH3OH–H2O solution. The compound crystallizes as heptahydrate. The crystal structure (monoclinic, P21/c (no. 14), a = 9.897(1), b = 26.384(2), c = 10.582(1) Å, β = 105.87(1)°, Z = 4, R = 0.0505, wR2 = 0.1029 for 3166 observed reflections (F ≥ 2σ(F ) out of 6100 unique reflections) consists of hydrogen bonded water molecules and [Ni(phen)2(CO3)] complex molecules. The Ni atoms are sixfold octahedrally coordinated by the four N atoms of two bidentate chelating phen ligands and by two O atoms of the bidentate chelating carbonate group with d(Ni–N) = 2.092–2.100 Å, d(Ni–O) = 2.051, 2.079 Å. The complex molecules are stacked into 2D corrugated layers parallel to (010) via two types of intermolecular π‐π stacking interactions. One occurs between two quinoline rings of neighboring phen ligands at the distance of 3.63 Å in [010] direction and the other results from 1D π‐π stacking interactions through partially covered phen rings at alternative distances of 3.26 Å and 3.33 Å in [001] direction. The water molecules are sandwiched between 2D layers.  相似文献   

15.
Chemically activated CF3SH, CFCl2SH, and CF2ClSH were formed through combination of SH and CF3, CFCl2, and CF2Cl radicals, respectively. The SH radical was prepared by abstraction of an H‐atom from H2S by the halocarbon radical produced during photolysis of (CF3)2C=O, (CFCl2)2C=O, or (CF2Cl)2C=O. 1,2‐HX (X = F, Cl) elimination reactions were observed from CF3SH, CFCl2SH, and CF2ClSH with products detected by GC‐MS. The combination reaction of CF2Cl radicals with SH radicals prepared CF2ClSH molecules with approximately 318 kJ/mol of internal energy. The experimental rate constants for elimination of HCl and HF from CF2ClSH were 3 ± 3 × 1010 and 2 ± 1 × 109 s?1, respectively. Comparison to Rice–Ramsperger–Kassel–Marcus (RRKM) calculated rate constants assigned the threshold energies as 171 ± 12 and 205 ± 12 kJ/mol for the unimolecular elimination of HCl and HF, respectively. Theoretical calculations using the B3PW91, MP2, and M062X methods with the 6311+G(2d,p) and 6‐31G(d',p') basis sets established that for a specific method the threshold energies differ by only 4 kJ/mol between the two different basis sets. There was wide variation among the three methods, but the M062X approach appeared to give threshold energies closest to the experimental values. Chemically activated CF3SH and CFCl2SH were also prepared with about 318 kcal mol?1 of internal energy, and the HX (X = F, Cl) elimination reactions were observed. Only HCl loss was detected from CFCl2SH, but the rate was too fast to measure with our kinetic method; however, based on our detection limit the HF elimination channel is at least 50 times slower.  相似文献   

16.
Smog chamber/FTIR techniques were used to measure k(Cl + HCF2OCF2OCF2‐CF2OCF2H) = k(Cl + HCF2O(CF2O)n(CF2CF2O)mCF2H) = (5.0 ± 1.4) × 10?17 cm3 molecule?1 s?1 in 700 Torr of N2/O2 diluent at 296 ± 1 K. The Cl‐initiated atmospheric oxidation of HCF2OCF2OCF2CF2OCF2H and the sample of HCF2O(CF2O)n(CF2CF2O)mCF2H used in this work gave COF2 in molar yields of (476 ± 36)% and (859 ± 63)%, respectively, with no other observable carbon containing products (i.e., essentially complete conversion of both hydrofluoropolyethers into COF2). The results are discussed with respect to the atmospheric chemistry and environmental impact of hydrofluoropolyethers of the general formula HCF2O(CF2O)n(CF2CF2O)mCF2H. © 2008 Wiley Periodicals, Inc. Int J Chem Kinet 40: 819–825, 2008  相似文献   

17.
NaZr2N2SCl: A Flux‐Stabilized Derivative of Zirconium(IV) Nitride Sulfide (Zr2N2S) The oxidation of zirconium metal with elemental sulfur and sodium azide (NaN3) should give access to zirconium(IV) nitride sulfide, Zr2N2S, which could crystallize isotypically with the trigonal rare‐earth(III) oxide sulfides M2O2S (M = Y, La–Lu). Appropriate molar admixtures of these reactants together with NaCl added as flux were heated for seven days at 850 °C in torch‐sealed evacuated silica tubes. As main product, however, pale yellow platelets with the composition NaZr2N2SCl (trigonal, R 3 m; a = 363.56(3), c = 2951.2(4) pm; Z = 3) emerged as single crystals. This pseudo‐quaternary compound crystallizes isotypically with e. g. LixEr2HyCl2 (x ≤ 1, y ≤ 2) in a (doubly) stuffed ZrBr‐type structure and contains at least structural domains of the hypothetical Ce2O2S‐analogous Zr2N2S. Zr4+ resides in monocapped trigonal anti‐prismatic sevenfold coordination of the anions (d(Zr–N) = 218 (3 ×) and 220 pm (1 ×), d(Zr–S/Cl) = 266 pm, 3 ×). Closest packed double‐layers of Zr4+ with all tetrahedral interstices occupied with N3– are sandwiched by layers of isoelectronic S2– and Cl anions. These anionic six‐layer slabs (S/Cl–Zr–N–N–Zr–S/Cl) pile up parallel (001) in a cubic closest packed fashion. Charge balance and structural consistence occurs between these layers by intercalation of Na+ within octahedral voids (d(Na–S/Cl) = 282 pm, 6 ×) of double‐layers of the indistinguishable heavy anions (S2– and Cl).  相似文献   

18.
The thermal decomposition of trifluoromethoxycarbonyl peroxy nitrate, CF3OC(O)O2NO2, has been studied between 278 and 306 K at 270 mbar total pressure using He as a diluent gas. The pressure dependence of the reaction was also studied at 292 K between 1.2 and 270 mbar total pressure. The rate constant reaches its high‐pressure limit at 70 mbar. The first step of the decomposition leads to CF3OC(O)O2 and NO2 formation, that is, CF3OC(O)O2NO2 + M ? CF3OC(O)O2 + NO2 + M (k1, k?1). Reaction (?1) was prevented by adding an excess of NO that reacts with the peroxy radical intermediate and leads to carbonyl fluoride (CF2O), carbon dioxide (CO2), nitrogen dioxide (NO2), and small quantities of CF3OC(O)O2C(O)OCF3. The kinetics of reaction (1) was determined by following the loss of CF3OC(O)O2NO2 via IR spectroscopy. The temperature dependence of the decomposition follows the equation k1(T) = 1.0 × 1016 e?((111±3)/(RT)) for the exponential term expressed in kJ mol?1. The values obtained for the kinetic parameters such as k1 at 298 K, the activation energy (Ea), and the preexponential factor (A) are compared with literature data for other acyl peroxy nitrates. The atmospheric thermal stability of CF3OC(O)O2NO2 and its dependence with altitude is discussed. © 2008 Wiley Periodicals, Inc. Int J Chem Kinet 40: 831–838, 2008  相似文献   

19.
The adiabatic mechanism of the reaction of trichloroethylene with O(3P), exploring the various O-atom addition and H-atom abstraction channels, is theoretically studied at the MP2/6-311++G(2d, 2p), MP2/aug-cc-pVTZ, CCSD/6-31G(d), G3, and CBS-QB3 levels of theory. From a kinetic point of view, the addition to the less substituted carbon atom of the double bond is more favorable than the addition to the more substituted carbon. Such O-atom addition reactions are favored over the one possible hydrogen-abstraction reaction. Calculations of the present study showed that five products are obtained: HCCl + C(O)Cl2 (P1), Cl + ClC(O)CHCl (P2), H + ClC(O)CCl2 (P3), Cl + HC(O)CCl2 (P4), and CH(O)Cl + CCl2 (P5). The products P2 and P4 are found to be the most favored ones. The kinetic calculations of rate constant in the range of 285–395 K are performed at the CBS-QB3 level of theory and are in conformity with the experimental outcomes.  相似文献   

20.
Electronically excited carbon dioxide (CO2*) is known for its broadband emission, and its detection can lead to valuable information; however, owing to its broadband characteristics, CO2* is difficult to isolate experimentally, and its chemical kinetics are not well known. Although numerous works have monitored CO2* chemiluminescence, a full kinetic scheme for the excited species has yet to be developed. To this end, a series of shock‐tube experiments was performed in H2‐N2O‐CO mixtures highly diluted in argon at conditions where emission from CO2* could be isolated and monitored. These results were used to evaluate the kinetics of CO2*, in particular the main CO2* formation reaction CO + O + M CO2* + M (R1). Based on collision theory, the quenching chemistry of CO2* was estimated for 11 collision partners. The final mechanism developed for CO2* consists of 14 reactions and 13 species. The rate for (R1) was determined to within about ±60% using low‐pressure experiments performed in five different (H2‐)N2O‐CO‐Ar mixtures, as follows: where R is the universal gas constant in cal/mol‐K and T is the temperature in K. Final mechanism predictions were compared with experiments at low and high pressures, with good agreement at both conditions for the temperature dependence of the peak CO2* and the CO2* species time histories. Comparisons were also made with previous experiments in methane–oxygen mixtures, where there was slight overprediction of CO2* experimental trends, but with the results otherwise showing a dramatic improvement over an earlier mechanism. Experimental results and model predictions were also compared with past literature rates for CO2*, with good agreement for peak CO2* trends and slight discrepancies in CO2* species time histories. Overall, the ability of the CO2* mechanism developed in this work to reproduce a range of experimental trends represents an important improvement over the existing knowledge base on chemiluminescence 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