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 234 毫秒
1.
Gas Phase Structure of CF3NCl2 and Preparation of CF3NCl2F+MF6? (M = As, Sb) and CF2 = NCl2F+SbF6? The gas phase structure of CF3NCl2 is reported. The following skeletal parameters are derived (ra-values, error limits are 3σ values): N? C = 1.470(6) Å, N? Cl = 1.733(3) Å, ClNCl = 111.5(4)° and ClNC = 107.6(5)°. CF3NCl2F+MF6? is prepared by fluorination of CF3NCl2 with XeF+MF6?. The same educt CF3NCl2 reacts with XeF+SbF6? at ?40°C to CF2 = NClF+SbF6? under elimination of ClF.  相似文献   

2.
The C,N-chelated tri-, di- and monoorganotin(IV) halides react with equimolar amounts of CF3COOAg to give corresponding C,N-chelated organotin(IV) trifluoroacetates. The set of prepared tri-, di- and monoorganotin(IV) trifluoroacetates bearing the LCN ligand (where LCN is 2-(N,N-dimethylaminomethyl)phenyl-) was structurally characterized by X-ray diffraction analyses, multinuclear NMR and IR spectroscopy. In the case of triorganotin(IV) trifluoroacetates and (LCN)2Sn(OC(O)CF3)2, no tendency to form hydrolytic products, or instability towards the moisture was observed. LCNRSn(OC(O)CF3)2 (where R is n-Bu or Ph) and LCNSn(OC(O)CF3)3 forms upon crystallization from THF in the air mainly dinuclear complexes in which the two tin atoms are interconnected either by hydroxo-bridges or by an oxo-bridge and/or by a bridging trifluoroacetate(s). In the case of hydrolysis of LCN(n-Bu)Sn(OC(O)CF3)2, a zwitterionic stannate of formula LCN(n-Bu)Sn(OC(O)CF3)2·CF3COOH was isolated from the mother liquor, too. Products of hydrolysis of LCN(n-Bu)Sn(OC(O)CF3)2 and LCNSn(OC(O)CF3)3, and some other oxygen bridged organotin(IV) compounds containing the same ligand, were tested as possible catalysts of some transesterification reactions as well as in direct dimethyl carbonate (DMC) synthesis from CO2 and methanol.  相似文献   

3.
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   

4.
Some reactions of amines with a perfluorovinylsulfide, (CF3)2CFSC(CF3)CFCF(CF3)2, prepared from hexafluoropropene (HFP) and sulfur [1] are compared to reactions of the same amines with the thermodynamic dimer of HFP, (CF3)2CCFCF2CF3. Many new ketenimines, eneamines, amidines, nitriles, and quinoline derivatives are reported.  相似文献   

5.
Treatment of bis(fluoroalkyl) phosphites (RFCH2O)2P(O)H, where RF was CF3 or C2F5 with sulfur in pyridine at 80 °C gave salts of structure [(RFCH2O)2P(O)SH]NC5H5 in 90 and 88% yield, respectively. The salts reacted with alkyl iodides in acetonitrile at 50 °C to furnish bis(fluoroalkyl) S-alkyl phosphorothiolates (RFCH2O)2P(O)SR, where R was Me, Et, n- and i-Pr (when RF = CF3) and Me (when RF = C2F5). Yields ranged from 21 to 57%. Bis(trifluoroethyl) S-methyl phosphorothiolate (CF3CH2O)2P(O)SMe underwent fluorination by silver(I) fluoride in acetonitrile at room temperature to yield the phosphorofluoridate (CF3CH2O)2P(O)F in 75% yield. Tris(fluoroalkyl) phosphorothionates (RFCH2O)3P = S, where RF was CF3, C2F5 and C3F7, were prepared in 30-34% yield by heating the tris(fluoroalkyl) phosphites (RFCH2O)3P and sulfur to 200 °C in a sealed tube for 8 h.  相似文献   

6.
N-(Trifluoromethylsulfonylimino)di-and-trifluoromethanesulfinimidoyl chlorides RS(=NSO2CF3)Cl (R = CF3, CHF2) react with potassium fluoride in 1,2-dimethoxyethane with formation of the corresponding N,N′-bis(trifluoromethylsulfonyl)fluoromethanesulfinimidamide potassium salts RS(=NSO2CF3)NSO2CF3 ?K. Analogous methanesulfinimidoyl and fluoromethanesulfinimidoyl chlorides (R = CH3, CH2F) fail to react with KF under similar conditions. Treatment of trifluoromethanesulfenamides CF3SNR2 with N,N-dichlorotrifluoromethanesulfonamide CF3SO2NCl2 leads to N,N-disubstituted N′-(trifluoromethylsulfonyl)trifluoromethanesul-finimidamides CF3S(=NSO2CF3)NR2. The reaction of N,N-dimethyl-N′-(trifluoromethylsulfonyl)trifluoromethanesulfinimidamide (R = CH3) with gaseous hydrogen chloride in diethyl ether gives sulfinimidoyl chloride CF3S(=NSO2CF3)Cl which could not be obtained by imination of CF3SCl.  相似文献   

7.
Perfluoroalkenyl phosphonates were formed along with Me3SiF using CF3CF=CF2, CF3CH=CF2, F5SCF=CF2 or F5SCH=CF2 and silylated phosphites, (R1O)2POSiMe3 (R1=Et, SiMe3). This straightforward method could be extended to perfluorobutadienes CF2=C(RF)C(RF)=CF2 (RF F=F, CF3). The formation of CF3C(=O)P(=O)(OSiMe3)2 and further reactions to yield bisphosphonates will be described. Acetylphosphonates, R2C(=O)P(=O)(OSiMe3)2 (R2=CH3, CF3) reacted with the ketimine, CH3C(=NiPr)Ph to give α-hydroxy-γ-imino phosphonates. Trifluoroacetylphenol and 2,6-bis(trifluoracetyl)-4-methyl-phenol have been proven to be versatile precursors for α-and γ-hydroxy phosphonates. Intermediates in these reactions were found to be cyclic λ5σ5P species.  相似文献   

8.
In attempts to obtain kinetic and mechanistic data required for an assessment of atmospheric fate of alternative halocarbons containing a CF3 group, reactions of the key free radical intermediates CF3OO and CF3O with several atmospheric compounds (i.e., NO, NO2, alkanes and alkenes) have been studied at 297 ± 2 K in 700 torr of air. Experiments employed the long path-FTIR spectroscopic method for product analysis and the visible (400 nm) photolysis of CF3NO → CF3 + NO as a source for the precursor radical CF3. Numerous labile and stable F-containing molecular products have been characterized based on kinetic and spectroscopic data obtained at sufficiently short photolysis time (≤1 min) to minimize heterogeneous decay on the reactor walls. Major new findings have been made for the reactions involving CF3O radicals. The behavior of CF3O radicals has been shown to be markedly different from that of CH3O radicals, i.e., (1) O2-reaction: no evidence for the F-atom transfer reaction CF3O + O2 → CF2 O + FOO; (2) NO-reaction: addition reaction CH3O + NO (+M) → CH3ONO (+M), but F-transfer reaction CF3O + NO → CF2O + FNO; (3) NO2-reaction: addition reaction for both radicals, but F-transfer reaction CF3 + NO2 → CF2O + FNO2 to a minor extent; (4) alkane-reaction: much faster H-abstraction by CF3O, comparable to HO; (5) alkene-reaction: much faster addition reaction of CF3O, comparable to HO. These results are summarized in this paper.  相似文献   

9.
The well known fluorosulfonyldifluoroacetyl fluoride (I), FOCCF2SO2F (I) quantitatively formed from sulfur trioxide and TFE through the tetrafluoroethanesultone has been converted into the octafluoro- -5-iodo-3-oxapentanesulfonyl fluoride (II) ICF2CF2OCF2CF2SO2F (II) by the well known reaction (1) involving MF, iodine, TFE in aprotic solvents.The iodo compound (II) allowed us to obtain TFE telomers having both fluorosulfonyl and iodo as terminal groups.The said telomers have been easily converted into surfactants (III) through fluorination and vinyl derivatives (IV) by dehalogenation.CF3CF2(CF2CF2)nOCF2CF2SO3M (III)CF2CF(CF2CF2)nOCF2CF2SO2F (IV)  相似文献   

10.
Reaction of phosphorus trichloride with tert-butanol and fluoroalcohols gave bis(fluoroalkyl) phosphites (RFO)2P(O)H in 42-89% yield, where RF=HCF2CH2, H(CF2)2CH2, H(CF2)4CH2, CF3CH2, C2F5CH2, C3F7CH2, (CF3)2CH, (FCH2)2CH, CF3(CH3)2C, (CF3)2CH3C, CF3CH2CH2, C4F9CH2CH2 and C6F13CH2CH2. Treatment of these with chlorine in dichloromethane gave the bis(fluoroalkyl) phosphorochloridates (RFO)2P(O)Cl in 49-96% yield. The chloridate (CF3CH2O)2P(O)Cl was isolated in much lower yield from the interaction of thionyl chloride with bis(trifluoroethyl) phosphite. Heating the latter in dichloromethane with potassium fluoride and a catalytic amount of trifluoroacetic acid gave the corresponding fluoridate (CF3CH2O)2P(O)F in 84% yield. Treatment of bis(trifluoroethyl) phosphite with bromine or iodine gave the bromidate (CF3CH2O)2P(O)Br and iodidate (CF3CH2O)2P(O)I in 51 and 46% yield, respectively. The iodidate is the first dialkyl phosphoroiodidate to have been isolated and characterised properly—its discovery lags behind the first isolation of a dialkyl phosphorochloridate by over 130 years. The fluoroalkyl phosphoryl compounds are generally more stable than known unfluorinated counterparts.  相似文献   

11.
The mass spectra of ten complexes derived from 3,3,3-trifluoroprop-1-yne, namely (CF3C?CH)CO2(CO)6, (CF3C?CH) [π? C5H5Ni]2, (CF3C?CH)3Co2(CO)4, CF3CH2C[Co(CO)3]3, are discussed in terms of their structures. The major processes observed can be satisfactorily explained in terms of (i) loss of carbonyl groups, if present; (ii) loss of one fluorine atom from the parention (iii) elimination of neutral metal fluorides, with ligand transfer reactions in the case of the iron complexes; (iv) elimination of neutral fluorocarbon molecules.  相似文献   

12.
The fundamental IR vibrational modes of trifluoroacetyl fluoride CF3C(O)F and trifluoroacetyl chloride CF3C(O)Cl have been re-examined by ab initio molecular orbital calculations and compared with literature assignments. Several bands of the IR spectrum are reassigned. The Q-branch and integrated absorption cross-sections have been measured for ν1, ν3, ν4 and ν11 fundamental bands for both pressurized and unpressurized samples on each molecule. The UV absorption spectra of CF3C(O)F and CF3C(O)Cl show a structureless continuum with a maximum at 21Onm (σmax=3.20±0.02 × 10−20 cm2 molecule−1) and 255 nm (σmax=7.66±0.26 × 10−20 cm2 molecule−1), respectively. The nature of the electronic transition giving rise to the UV absorption spectrum for CF3C(O)F and CF3C(O)Cl has been examined by ab initio molecular orbital calculations. It is attributed to the A1A″←X1A′ electronic transition.  相似文献   

13.
The unimolecular decomposition reaction of CF3CCl2O radical has been investigated using theoretical methods. Two most important channels of decomposition occurring via C–C bond scission and Cl elimination have been considered during the present investigation. Ab initio quantum mechanical calculations are performed to get optimized structure and vibrational frequencies at DFT and MP2 levels of theory. Energetics are further refined by the application of a modified Gaussian-2 method, G2M(CC,MP2). The thermal rate constants for the decomposition reactions involved are evaluated using Canonical Transition State Theory (CTST) utilizing the ab initio data. Rate constants for C–C bond scission and Cl elimination are found to be 6.7 × 106 and 1.1 × 108 s?1, respectively, at 298 K and 1 atm pressure with an energy barrier of 8.6 and 6.5 kcal/mol, respectively. These values suggest that Cl elimination is the dominant process during the decomposition of the CF3CCl2O radical. Transition states are searched on the potential energy surface of the decomposition reactions involved and are characterized by the existence of only one imaginary frequency (NIMAG = 1) during frequency calculation. The existence of transition states on the corresponding potential energy surface is further ascertained by performing intrinsic reaction coordinate (IRC) calculation.  相似文献   

14.
The synthesis and X-ray single crystal study of two mixed-ligand Cu(II) complexes are performed: (CH3C(NCH3)CHC(O)CH3)(CF3C(O)CHC(O)CF3)Cu (1) (space group P21/c, a = 7.0848(12) Å, b = 17.854(3) Å, c = 11.837(2) Å, β = 100.495(6)°, V = 1472.4(4) Å3, Z = 4), (CH3C(NC6H5)CHC(O)CH3)· (CF3C(O)CHC(O)CF3)Cu (2) (space group P-1, a = 9.1119(4) Å, b = 9.6954(4) Å, c = 11.1447(6) Å, α = 113.784(2)°, β = 92.383(2)°, γ = 95.402(2)°, V = 893.52(7) Å3, Z = 2). The structures are molecular, formed from neutral mixed-ligand copper complexes. The central copper atom has the (3O+N) coordination environment with average Cu-O distances of 1.948 Å and Cu-N of 1.932 Å; the chelate O-Cu-N angle (average) is 94.0°. In the structures, the complexes are linked into dimeric associates with Cu…Cu distances of 3.197 Å (for 1) and 3.246 Å (for 2). The volatility of mixed-ligand complexes 1 and 2 is in between of that of the starting homo-ligand complexes.  相似文献   

15.
Bis-alkenyl complexes of the type (η-C5H5)2RH2(alkene − H)(alkyne + H) are obtained when the alkyne complex (η-C5H5)2Rh2(CO)(CF3C2CF3) is treated with the following alkenes: H2CCH2, H2CCHR (R = Me, But, Ph, CN), H2CCF2, RHCCHR′ (R = R′ = Me, Ph, Cl; R = Me, R′ = Et), cyclooctene and norbornene. An approximately equimolar amount of (η-C5H5)2Rh2(CO)2(CF3C2CF3) is also formed. The reactions are greatly accelerated when the reaction mixtures are exposed to sunlight. There is some regioselectivity in the reactions with H2CCHR and MeHCCHet, with a preference for CH bond cleavage at the least crowded alkene-carbon. When the reaction with acrylonitrile is performed in the absence of sunlight, the complex (η-C5H5)2(CO){(H2CCHCN)(CF3C2CF3)} can be isolated; upon exposure to sunlight, there is loss of CO and H-transfer to form two isomers of the appropriate bis-alkenyl complex.The molecular geometries of (η-C5H5)2Rh2(CHCHCN){C(CF3)C(CF3)H} and (η-C5H5)2Rh2(CHCF2){C(CF3)C(CF3)H} have been ascertained by X-ray structure determination. Each molecule has two bridging alkenyl units spanning a RhRh single bond; the dihedral angle between the two RhRhCC planes is just above 90°. There is a cyclopentadienyl ring η5-attached to each metal. Crystal data: C17H13F6NRh2·H2O, M 569.1, monoclinic, P21/n, a 15.014(7), b 14.882(7), c 8.590(5) Å, β 94.57(9)°, Z = 4, final R 0.056 for 2493 observed reflections; C16H12F8Rh2, M 562.1, monoclinic, P21/c, a 13.037(6), b 8.765(2), c 14.873(3) Å, β 103.16(3)°, Z = 4, final R 0.062 for 1820 observed reflections.  相似文献   

16.
For many thermal reactions, the effects of catalysis or the influence of solvents on reaction rates can be rationalized by simple transition state models. This is not the case for reactions controlled by quantum tunneling, which do not proceed via transition states, and therefore lack the simple concept of transition state stabilization. 1H-Bicyclo[3.1.0]-hexa-3,5-dien-2-one is a highly strained cyclopropene that rearranges to 4-oxocyclohexa-2,5-dienylidene via heavy-atom tunneling. H2O, CF3I, or BF3 form Lewis acid–base complexes with both reactant and product, and the influence of these intermolecular complexes on the tunneling rates for this rearrangement was studied. The tunneling rate increases by a factor of 11 for the H2O complex, by 23 for the CF3I complex, and is too fast to be measured for the BF3 complex. These observations agree with quantum chemical calculations predicting a decrease in both barrier height and barrier width upon complexation with Lewis acids, resulting in the observed Lewis acid catalysis of the tunneling rearrangement.

The ring-opening of a highly strained cyclopropene to a carbene proceeds via heavy-atom tunneling. This rearrangement is accelerated in the presence of H2O, ICF3 or BF3, resulting in a novel Lewis-acid catalyzed tunneling reaction.  相似文献   

17.
Data on the tropospheric degradation of proposed substitutes for ozone depleting CFCs were obtained by conducting photochemical oxidation studies of HCFCs and HFCs using long path Fourier transform infrared spectroscopy. The hydrogen abstraction reactions were initiated using Cl radicals rather than OH radicals because of the rather unreactive nature of the compounds. The experimental product yields at T = 25 ± 3°C and 700 Torr of dry air were: CHClF2 (1.11 ± 0.06 C(O)F2); CClFHCF3 (1.00 ± 0.04 CF3C(O)F); CF3CHF2 (1.09 ± 0.05 C(O)F2); CClF2CH3 (0.98 ± 0.03 C(O)F2); CHF2CH3 (1.00 ± 0.05 C(O)F2); CF3CH2F (0.16 ± 0.03 CF3CF(O), and 0.83 ± 0.22 HFC(O)), where all standard deviations are 2σ. For each compound, the critical step in determining the oxidation products was the decomposition of a halogenated alkoxy radical. For HCFC-22 and HCFC-124, the major alkoxy radical decomposition route was Cl elimination. The HFC-125 product data were consistent with C? C cleavage of a two carbon alkoxy radical as the major decomposition route whereas both C? C cleavage and H abstraction by O2 were significant contributors to the decomposition of the HFC-134a alkoxy radical. Secondary Cl reactions in the HCFC-142b and HFC-152a experiments prevented an unambiguous determination of the decomposition modes; the data are consistent with both C? C bond scission and Cl reactions with halogenated aldehydes producing the oxidation product C(O)F2. With the exception of the HFC-134a and HFC-125 data, the proposed mechanisms can account for the major oxidation products. For HFC-134a and HFC-125, a number of product bands could not be identified. The bands are likely due to products from reactions involving the CF3O2 radical. © John Wiley & Sons, Inc.  相似文献   

18.
Bis(cyclopentadienyl)mercury readily undergoes Diels—Alder reactions with RCCR (R = CO2Me or CF3), CF3CFCFCF3, CF3CFCF2, (CF3)2CC(CN)2, C2(CN)4 and PhNCONNCO to give stable adducts characterised by1H, 19F and 13C NMR, spectroscopy. Similar reactions of CF3CCCF3 and CF3CFCFCF3 with the cyclopentadiene derivatives Me3MC5H5 and (Me3M)2C5H4 (M = Si, Sn) are also described.  相似文献   

19.
The reactivity of bis(fluoroalkyl) phosphorochloridates to nucleophiles is summarised. Previous data and the results described here indicate that reactivities decrease in the order: amines>alcohols>thiols. The synthesis of CF3CH2OP(O)(SEt)2 in 30% yield was accomplished by treating CF3CH2OP(O)Cl2 with two molar equivalents of EtSH and Et3N in ether. The chloridates (CF3CH2O)2P(O)Cl and (C2F5CH2O)2P(O)Cl did not react with MeSH in ether at −78 °C or when heated with Pb(SMe)2 in benzene. Ethanethiol and propanethiol reacted with fluorinated chloridates in the presence of triethylamine to give thiolates (RFO)2P(O)SR in 13-41% yield where RF was CF3CH2, C2F5CH2, C3F7CH2 or (CF3)2CH and R was Et or n-Pr. Similarly, reaction of phosphorobromidates (RFCH2O)2P(O)Br, made by brominating the corresponding bis(fluoroalkyl) H-phosphonates, with benzenethiol gave derivatives (RFCH2O)2P(O)SPh in 43 and 46% yield where RF was CF3 and C2F5, respectively. Treatment of the chloridothiolate Cl(EtO)P(O)SMe, prepared in two steps from triethyl phosphite, with fluoroalcohols and triethylamine in ether gave species RFO(EtO)P(O)SMe in 62-74% yield where RF was CF3CH2, C2F5CH2, C3F7CH2 or (CF3)2CH. The reactions of bis(trifluoroethyl) phosphorochloridate with 2-mercaptoethanol, 3-mercaptopropanol and ethane-1,2-dithiol gave several unexpected products whose structures were tentatively assigned.  相似文献   

20.
The mechanism, kinetics, and thermochemistry of the gas-phase reactions of CF2ClC(O)OCH2CH3,ethyl chlorodifluoroacetate (ECDFA) with the OH radical and Cl atom are investigated. Geometry optimization and frequency calculations have been performed at the MPWB1K/6-31+G(d,p) level of theory and energetic information is refined by using G2(MP2) theory. Transition states are searched on the potential energy surface of reaction channels and each of the transition states is characterized by the presence of only one imaginary frequency. Connections of the transition states between designated local minima are confirmed by intrinsic reaction coordinate calculation. Theoretically calculated rate constants at 298 K using the Canonical Transition State Theory are found to be in good agreement with the experimentally measured ones. Using group-balanced isodesmic reactions as working chemical reactions, the standard enthalpies of formation for CF2ClC(O)OCH2CH3, CF2ClC(O)OCH2CH2, and CF3C(O)OCHCH3 are also reported for the first time. The hydrogen abstraction occurs mainly from –CH2 group. The T1 diagnostic calculation suggests that the multi-reference character is not an issue for such systems. The estimated atmospheric life time of ECDFA is expected to be around 24 days.  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