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 46 毫秒
1.
The atmospheric chemistry of CCl2FCH2CF3 (HFCF-234fb) was examined using FT-IR/relative-rate methods. Hydroxyl radical and chlorine atom rate coefficients of k(CCl2FCH2CF3+OH)= (2.9 ± 0.8) × 10−15 cm3 molecule–1 s–1 and k(CCl2FCH2CF3+Cl)= (2.3 ± 0.6) × 10−17 cm3 molecule–1 s–1 were determined at 297 ± 2 K. The OH rate coefficient determined here is two times higher than the previous literature value. The atmospheric lifetime for CCl2FCH2CF3 with respect to reaction with OH radicals is approximately 21 years using the OH rate coefficient determined in this work, estimated Arrhenius parameters and scaling it to the atmospheric lifetime of CH3CCl3. The chlorine atom initiated oxidation of CCl2FCH2CF3 gives C(O)F2 and C(O)ClF as stable secondary products. The halogenated carbon balance is close to 80% in our system. The integrated IR absorption cross-section for CCl2FCH2CF3 is 1.87 × 10−16 cm molecule−1 (600–1600 cm−1) and the radiative efficiency was calculated to 0.26 W m−2 ppb1. A 100-year Global Warming Potential (GWP) of 1460 was determined, accounting for an estimated stratospheric lifetime of 58 years and using a lifetime-corrected radiative efficiency estimation.  相似文献   

2.
Phosphonium Salts with Hydrogen Dihalide Anions HCl2?, HBr2?, HI2?, or HBrCl? Phosphonium hydrogen dihalides [R3PR′][XHY] (X = Y = Cl, Br, I; X = Br, Y = Cl) resp. [R3PH]HBr2 are obtained as extremely hydrolyzable crystals by reaction of phosphonium halides or tertiary phosphanes with hydrogen halide. According to IR spectroscopic results the solid compounds mostly contain anions [XHX]? with symmetric hydrogen bonds. In solution 1H NMR measurements show a slight (X = Cl, Br) or considerable (X = I) dissociation according to HX2? ? X? + HX. On heating the solid compounds decompose with formation of hydrogen halide and [R3PR′]X or [R3PH]X. In this process the hydrogen bromidechlorides [R3PR′][BrHCl] exclusively eliminate HCl. NMR studies (1H und 31P) with solutions containing [R3PH]HBr2 (R = phenyl, 1-naphtyl) or HBr and Ph3P in varying molar ratios show that a fast proton exchange between the competing Lewis bases R3P and Br? exists.  相似文献   

3.
In this study, we report the synthesis and reactivity of [18F]fluoromethyl iodide ([18F]FCH2I) with various nucleophilic substrates and the stabilities of [18F]fluoromethylated compounds. [18F]FCH2I was prepared by reacting diiodomethane (CH2I2) with [18F]KF, and purified by distillation in radiochemical yields of 14-31% (n = 25). [18F]FCH2I was stable in organic solvents commonly used for labeling and aqueous solution with pH 1-7, but was unstable in basic solutions. [18F]FCH2I displayed a high reactivity with various nucleophilic substrates such as phenol, thiophenol, amide and amine. The [18F]fluoromethylated compounds synthesized by the reactions of phenol, thiophenol and tertiary amine with [18F]FCH2I were stable for purification, formulation and storage. In contrast, the [18F]fluoromethylated compounds synthesized by the reactions of primary or secondary amines, and amide with [18F]FCH2I were too unstable to be detected or purified from the reaction mixtures. Defluorination of these [18F]fluoromethyl compounds was a main decomposition route.  相似文献   

4.
The determination of minima and saddle points on the potential energy surfaces of the hydrogen bonded species O2?HF and O2?H2O is performed with unrestricted Hartree-Fock calculations. Geometries, electron density distributions, and relative energies for every stationary point are reported. Only one true minimum is found for O2?HF and for O2?H2O, and this approximately corresponds to a structure where the partially positive hydrogen atom is located along one of the superoxide ion electron lone-pair directions. Calculated ΔH, ΔS, and ΔG values for the reaction between O2? and H2O are in good agreement with experimental data.  相似文献   

5.
Multiple bonds between boron and transition metals are known in many borylene (:BR) complexes via metal dπ→BR back‐donation, despite the electron deficiency of boron. An electron‐precise metal–boron triple bond was first observed in BiB2O? [Bi≡B?B≡O]? in which both boron atoms can be viewed as sp‐hybridized and the [B?BO]? fragment is isoelectronic to a carbyne (CR). To search for the first electron‐precise transition‐metal‐boron triple‐bond species, we have produced IrB2O? and ReB2O? and investigated them by photoelectron spectroscopy and quantum‐chemical calculations. The results allow to elucidate the structures and bonding in the two clusters. We find IrB2O? has a closed‐shell bent structure (Cs, 1A′) with BO? coordinated to an Ir≡B unit, (?OB)Ir≡B, whereas ReB2O? is linear (C∞v, 3Σ?) with an electron‐precise Re≡B triple bond, [Re≡B?B≡O]?. The results suggest the intriguing possibility of synthesizing compounds with electron‐precise M≡B triple bonds analogous to classical carbyne systems.  相似文献   

6.
Collision-induced dissociation of the ions [ArS]?, [ArSO]? and [ArSO2]? has uncovered a rich and varied ion chemistry. The major fragmentations of [ArS]? are complex and occur without prior ring hydrogen scrambling: for example, [C6H5S]?→[C2HS]? and [HS]?; [p-CD3C6H4S]?→[C6H4S]?˙, [CD3C4S]? and [C2HS]?. In contrast, all decompositions of [C6H5CH2S]? are preceded by specific benzylic and phenyl hydrogen interchange reactions. [ArSO2]? and [ArSO2]? ions undergo rearrangement, e.g. [C6H5SO]?→[C6H5O]? and [C6H5S]?; [C6H5SO2]?→[C6H5O] ?. The ion [C6H5CH2SO]? eliminates water, this decomposition is preceded by benzylic and phenyl hydrogen exchange.  相似文献   

7.
Gas phase reactions between PtHn? cluster anions and CO2 were investigated by mass spectrometry, anion photoelectron spectroscopy, and computations. Two major products, PtCO2H? and PtCO2H3?, were observed. The atomic connectivity in PtCO2H? can be depicted as HPtCO2?, where the platinum atom is bonded to a bent CO2 moiety on one side and a hydrogen atom on the other. The atomic connectivity of PtCO2H3? can be described as H2Pt(HCO2)?, where the platinum atom is bound to a formate moiety on one side and two hydrogen atoms on the other. Computational studies of the reaction pathway revealed that the hydrogenation of CO2 by PtH3? is highly energetically favorable.  相似文献   

8.
Preparation of μ-Sulfurdisulfonium Salts [(CH3)2S? Sx? S(CH3)2]2+2A? (x = 1–3, A? = AsF6?, SbF6?, SbCl6?). On the Analogy of the Reactivity of Sulfanes and Sulfonium Salts The preparation of the μ-sulfurdisulfonium salts [(CH3)2S? Sx? S(CH3)2]2+(A?)2 with x = 1–3 and A? = AsF6?, SbF6?, SbCl6? is reported. The salts are formed by reaction of (CH3)2SH+A? and (CH3)2SSH+A? with SCl2 and S2Cl2, resp. They are characterized by vibrational spectroscopic measurements. [(CH3)2S? S2? S(CH3)2]2+(SbF6?)2 crystallizes in the space group C2/c with a = 1 884.5(7) pm, b = 1 302.8(5) pm, c = 1 477.2(5) pm, β = 98.62(3)° und Z = 8.  相似文献   

9.
A procedure to calculate the quantum mechanical transition probability of a unimolecular primary chemical process, A?A + e? is investigated for the circumstance where A? and A have different numbers of vibrational and rotational degrees of freedom (one is linear, the other not). A procedure is introduced to deal with the coupling between the vibrational and rotational motions. The proposed method was applied to calculating the lifetimes of CO2˙? and N2O˙? in the gas phase. The geometry optimizations and frequency calculations for CO2, CO2˙?, N2O, and N2O˙? are performed at HF, MP2, and QCISD(T) levels with 6-31G* or 6–31+G* basis sets, in order to obtain reliable geometric and spectroscopic information on these systems. Lifetimes are calculated for several of the lower vibrational–rotational states of the anions, as well as for the Boltzmann distribution of states at 298 K. The lifetime of the lowest vibrational–rotational state of CO2˙?, is 1.03 × 10?4 s, and of the lowest vibrational state with rotational levels weighted by Boltzmann distribution at 298 K, 1.50 × 10?4 s. These values are in good agreement with the experimental number, 9.0 ± 2.0 × 10?5 s, and support the experimental evidence that CO2˙? was formed in its ground vibrational level by the techniques used. The lifetime of CO2˙? calculated with Boltzmann distribution over its vibrational and rotational levels at 298 K, is 1.51 × 10?5 s. There are no direct measurements of the lifetime of N2O˙?, but it was estimated to be greater than 10?4 s from experimental evidence. The predicted lifetimes of N2O˙?, at its lowest vibrational–rotational state (0 K) and lowest vibrational state with rotational levels weighted by the Boltzmann distribution at 298 K, are 238 and 19.1 s, respectively. The lifetime of N2O˙? at thermal equilibrium at 298 K is 6.66 × 10?2 s, indicating that electron loss from the excited vibrational states of N2O˙? is significant. This study represents the first theoretical investigation of CO2˙? and N2O˙? lifetimes. © 1994 John Wiley & Sons, Inc.  相似文献   

10.
Understanding and controlling the kinetics of O2 reduction in the presence of Li+‐containing aprotic solvents, to either Li+‐O2? by one‐electron reduction or Li2O2 by two‐electron reduction, is instrumental to enhance the discharge voltage and capacity of aprotic Li‐O2 batteries. Standard potentials of O2/Li+‐O2? and O2/O2? were experimentally measured and computed using a mixed cluster‐continuum model of ion solvation. Increasing combined solvation of Li+ and O2? was found to lower the coupling of Li+‐O2? and the difference between O2/Li+‐O2? and O2/O2? potentials. The solvation energy of Li+ trended with donor number (DN), and varied greater than that of O2? ions, which correlated with acceptor number (AN), explaining a previously reported correlation between Li+‐O2? solubility and DN. These results highlight the importance of the interplay between ion–solvent and ion–ion interactions for manipulating the energetics of intermediate species produced in aprotic metal–oxygen batteries.  相似文献   

11.
The interaction of superoxide ion O2? with up to four water molecules [O2?: (H2O)n, n = 1, 2, 4] has been investigated using ab initio molecular orbital theory. The binding energy of O2?: H2O is calculated to be ?20.6 kcal/mol in good agreement with gas phase experimental data. At the MP3/6-31G* level the O2?:H2O complex has a C2v structure with a double (cyclic) hydrogen bond between O2? and H2O. A Cs structure with a single hydrogen bond is only 0.7 kcal/mol less stable. Interaction of H2O with the doubly occupied π* orbital of O2? is preferred slightly over interaction with the singly occupied π* orbital. Natural bond orbital analysis suggests that both electrostatic and charge transfer interactions are important in anionic complexes. The charge transfer occurs predominantly in the O2? → H2O direction and is important in determining the relative stabilities of the different structures and states. Singly and doubly hydrogen-bonded structures for the O2?: (H2O)2 and O2?: (H2O)4 clusters were found to be similar in stability and the increase in binding of the cluster becomes smaller as each additional water molecule is added to the cluster.  相似文献   

12.
The structures and energies of the electronic ground states of the FeS0/?, FeS20/?, Fe2S20/?, Fe3S40/?, and Fe4S40/? neutral and anionic clusters have been computed systematically with nine computational methods in combination with seven basis sets. The computed adiabatic electronic affinities (AEA) have been compared with available experimental data. Most reasonable agreements between theory and experiment have been found for both hybrid B3LYP and B3PW91 functionals in conjugation with 6‐311+G* and QZVP basis sets. Detailed comparisons between the available experimental and computed AEA data at the B3LYP/6‐311+G* level identified the electronic ground state of 5Δ for FeS, 4Δ for FeS?, 5B2 for FeS2, 6A1 for FeS2?, 1A1 for Fe2S2, 8A′ for Fe2S2?, 5A′′ for Fe3S4, 6A′′ for Fe3S4?, 1A1 for Fe4S4, and 1A2 for Fe4S4?. In addition, Fe2S2, Fe3S4, Fe3S4?, Fe4S4, and Fe4S4? are antiferromagnetic at the B3LYP/6‐311+G* level. The magnetic properties are discussed on the basis of natural bond orbital analysis.  相似文献   

13.
The reaction thermodynamics of the 1,2‐dimethoxyethane (DME), a model solvent molecule commonly used in electrolytes for Li?O2 rechargeable batteries, has been studied by first‐principles methods to predict its degradation processes in highly oxidizing environments. In particular, the reactivity of DME towards the superoxide anion O2? in oxygen‐poor or oxygen‐rich environments is studied by density functional calculations. Solvation effects are considered by employing a self‐consistent reaction field in a continuum solvation model. The degradation of DME occurs through competitive thermodynamically driven reaction paths that end with the formation of partially oxidized final products such as formaldehyde and methoxyethene in oxygen‐poor environments and methyl oxalate, methyl formate, 1‐formate methyl acetate, methoxy ethanoic methanoic anhydride, and ethylene glycol diformate in oxygen‐rich environments. This chemical reactivity indirectly behaves as an electroactive parasitic process and therefore wastes part of the charge exchanged in Li?O2 cells upon discharge. This study is the first complete rationale to be reported about the degradation chemistry of DME due to direct interaction with O2?/O2 molecules. These findings pave the way for a rational development of new solvent molecules for Li?O2 electrolytes.  相似文献   

14.
The NCI(F?) and NCI(NH2?) mass spectra of a series of aliphatic acetates and of methyl and ethyl trimethylacetate have been obtained. The formation of fluoroenolate ions CH2COF? and of carboxamide anions RCONH? (R ? CH3))CH3C). respectively, is observed besides formation of [M ? H]? ions and carboxylate ions RCOO? (R ? CH3, (CH3)3C). The relative intensities of the different anions depend on the structure of the ester molecules and on the primary reactant anions. Usually, the NCI(NH2?) spectra of the acetates are dominated by [M ? H]? ions ([M? D]? ions in the case of trideuteroacetates) fragmenting unimolecularly by elimination of an alcohol. The carboxylate ions are important fragments, too, but carboxamide ions are only observed with large intensities in the NCI(NH2? spectra of the trimethylacetates. The NCI(F?) spectra show much larger intensities of carboxylate ions and fluoroenolate ions. The mechanisms of the fragmentation reactions are discussed. The results indicate that most or even all of the fragment ions in the NCI(F? mass spectra of aliphatic esters are formed by addition-elimination reactions via a tetrahedral intermediate, while competition between direct proton abstraction and addition-elimination reactions occurs in the NCI(NH2?) mass spectra because of the higher basicity of NH2? resulting in an early transition state for direct proton abstraction.  相似文献   

15.
The Synthesis of the Dichloromethylene-halogenosulfenium Salts Cl2CSCl+ AsF6? and Cl2CSBr+ AsF6? The sulfenium salts Cl2CSCl+ AsF6? and Cl2CSBr+ AsF6? are synthesized by oxidative halogenation of thiophosgene, Cl2CS with X2/AsF5 (X = Cl, Br) at 195 K and are characterized by vibrational as well as NMR spectroscopy.  相似文献   

16.
There are four neutralization-reionization processes which can be studied. In the first two, gaseous cations are collisionally reduced and the resulting fast neutrals are either oxidized by collisional electron removal (NR) or reduced further by collisional electron attachment (NR?). In the third and fourth processes, gaseous anions are collisionally oxidized and the resulting fast neutrals are either oxidized further by collisional electron removal (?NR) or reduced by collisional electron attachment (?NR?). These four processes are illustrated briefly using the interrelated species [SF5]+ and [SF5]?. Negative-ion-negative-ion neutralization-reionization (?NR?) processes are illustrated for [CH3O]?, [C2H5O]?, [CO3]? and [HCO3]? and compared with ?NR processes. In addition, collisional electron attachment ionization of fast neutrals formed in unimolecular fragmentation of cations is illustrated. The ?NR? studies of [CO3]? and [HCO3]? show that both CO3 and HCO3 are stable neutral species with lifetimes greater than 0.7 μs.  相似文献   

17.
Gas-phase reactions typical of the Earth’s atmosphere have been studied for a number of partially fluorinated alcohols (PFAs). The rate constants of the reactions of CF3CH2OH, CH2FCH2OH, and CHF2CH2OH with fluorine atoms have been determined by the relative measurement method. The rate constant for CF3CH2OH has been measured in the temperature range 258–358 K (k = (3.4 ± 2.0) × 1013exp(?E/RT) cm3 mol?1 s?1, where E = ?(1.5 ± 1.3) kJ/mol). The rate constants for CH2FCH2OH and CHF2CH2OH have been determined at room temperature to be (8.3 ± 2.9) × 1013 (T = 295 K) and (6.4 ± 0.6) × 1013 (T = 296 K) cm3 mol?1 s?1, respectively. The rate constants of the reactions between dioxygen and primary radicals resulting from PFA + F reactions have been determined by the relative measurement method. The reaction between O2 and the radicals of the general formula C2H2F3O (CF3CH2? and CF3?HOH) have been investigated in the temperature range 258–358 K to obtain k = (3.8 ± 2.0) × 108exp(?E/RT) cm3 mol?1 s?1, where E = ?(10.2 ± 1.5) kJ/mol. For the reaction between O2 and the radicals of the general formula C2H4FO (? HFCH2O, CH2F?HOH, and CH2FCH2?) at T = 258–358 K, k = (1.3 ± 0.6) × 1011exp(?E/RT) cm3 mol?1 s?1, where E = ?(5.3 ± 1.4) kJ/mol. The rate constant of the reaction between O2 and the radicals with the general formula C2H3F2O (?F2CH2O, CHF2?HOH, and CHF2CH2?) at T = 300 K is k = 1.32 × 1011 cm3 mol?1 s?1. For the reaction between NO and the primary radicals with the general formula C2H2F3O (CF3CH2? and CF3?HOH), which result from the reaction CF3CH2OH + F, the rate constant at 298 K is k = 9.7 × 109 cm3 mol?1 s?1. The experiments were carried out in a flow reactor, and the reaction mixture was analyzed mass-spectrometrically. A mechanism based on the results of our studies and on the literature data has been suggested for the atmospheric degradation of PFAs.  相似文献   

18.
Investigations on the reactivity of atomic clusters have led to the identification of the elementary steps involved in catalytic CO oxidation, a prototypical reaction in heterogeneous catalysis. The atomic oxygen species O.? and O2? bonded to early‐transition‐metal oxide clusters have been shown to oxidize CO. This study reports that when an Au2 dimer is incorporated within the cluster, the molecular oxygen species O22? bonded to vanadium can be activated to oxidize CO under thermal collision conditions. The gold dimer was doped into Au2VO4? cluster ions which then reacted with CO in an ion‐trap reactor to produce Au2VO3? and then Au2VO2?. The dynamic nature of gold in terms of electron storage and release promotes CO oxidation and O? O bond reduction. The oxidation of CO by atomic clusters in this study parallels similar behavior reported for the oxidation of CO by supported gold catalysts.  相似文献   

19.
Ab initio Hartree–Fock calculations are performed on hydrates of the F? and Cl? ions using 6-31G, 6-31G**, and 6-21G basis sets. Geometries and binding energies are obtained. An estimate of the correlation energy is provided by an MP2/6-31G (Møller-Plesset second-order perturbation) calculation. Comparisons are made between the Cl?(SO2) and the Cl?(H2O) complexes.  相似文献   

20.
Crystal Structures of SeCl3+SbCl6?, SeBr3+GaBr4?, PCl4+SeCl5?, and (PPh4+)2SeCl42? · 2 CH3CN The crystal structures of the title compounds were determined by X-ray diffraction. SeCl3+SbCl6?: Space group P21/m, Z = 4, structure determination with 1795 observed unique reflections, R = 0.022. Lattice dimensions at ?80°C: a = 940.9, b = 1066.3, c = 1234.9 pm, β = 102.79°. The compound forms ion pairs with the structure of a double octahedron with linked surfaces. SeBr3+GaBr4?: Space group Pc, Z = 2, structure determination with 1461 observed unique reflections, R = 0.058. Lattice dimensions at ?60°C: a = 660.1, b = 655.3, c = 1431.3 pm, β = 101.177°. The compound crystallizes in the SCl3[AlCl4] lattice type. Between the ions there are two relatively short Se … Br? Ga contacts. PCl4+SeCl5?: Space group Ima2, Z = 8, structure determination with 1757 observed unique reflections, R = 0.029. Lattice dimensions at ?50°C: a = 1651.6, b = 1201.2, c = 1166.4 pm. The SeCl5? ions are associated to chains via interionic Se? Cl … Se contacts along the crystallographic c-axis. (PPh4+)2SeCl42? · 2CH3CN: Space group P21/n, Z = 2, structure determination with 2578 observed unique reflections, R = 0.050. Lattice dimensions at ?80°C: a = 1288.5, b = 726.0, c = 2585.8 pm, β = 101.65°. The compound includes planar-tetragonal SeCl42? ions, which almost meet D4h symmetry.  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