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 15 毫秒
1.
Infrared spectra obtained for seven alkyl esters of pentafluoropropionic acid show that six of the nine stretching vibrations of the CF3CF2COO group and the three CF3 deformation vibrations have characteristic frequencies, with average deviations of 5 cm-1 or less. These frequencies are compared with the corresponding values in the acid and the trifluoroacetate esters.  相似文献   

2.
The behaviour of a series of differently halogenated acetic acid ethyl esters was studied using electron impact and metastable ion techniques. The presence of different halogens strongly influences the fragmentation pathways and also the kinetic energy release values associated with some decomposition routes. By mass-analysed ion kinetic energy spectrometry, a composite metastable peak was detected for the transition CF2BrCO+ → CF2Br+, and rationalized on the basis of two different structures for the precursor ion.  相似文献   

3.
The reaction of some diaryl-substituted olefins [1,1-diphenylethylene (DPE), 1,1,3,3-tetraphenyl 1-butene (TPB) and 3-phenylindene (PI)] with trifluoromethanesulphonic acid has been studied in dichloromethane by means of u.v. and 1H NMR spectroscopies. All carbocations are stable at temperatures below -30°. In the case of DPE at ?70°, the carbocation initially formed is dimeric but is slowly transformed into a monomeric cation. The initiation step is too fast to be studied by conventional means except for TPB where a first order rate with respect to the olefin has been observed.For DPE and TPB, when the ratio olefin/acid is high, a 33% yield in carbocations with respect to CF3SO3H is obtained for the whole range of acid concentration investigated (10?4 ?1 mol 1?1). In the case of PI, the yield is 50%. 1H NMR spectra clearly show that some acid remains unreacted in the presence of olefin. The large downfield shift of unconsumed acid is consistent with the existence of complex anions CF3SO?3, (HOSO2CF3)2 and CF3SO?3, HOSO2CF3 resulting from strong hydrogen bonds between the triflate anion and its conjugate acid.  相似文献   

4.
Several hydrated lanthanide salts of heptafluorobutyric acid have been prepared and characterized.Hydration studies have shown the compounds to exist in various hydration states across the series. Powder X-ray diffraction studies support the existence of variable hydration states in that three distinct crystalline forms exist. Infrared analysis revealed two types of structures to be present. Upon dehydration of the lighter rare earth hydrates to intermediate dihydrates, all of the compounds showed similar structures. Decomposition was found to be exothermic. The volatile decomposition products were identified by infrared analysis and found to consist of CO, CO2, CF3CF2COF and CF3CF2CF2COF. The amounts of each gas were found to be dependent on the decomposition temperature. The non-volatile decomposition products were identified by powder X-ray diffraction and found to consist of LnF3, LnOF and Ln2O3.  相似文献   

5.
In order to synthesize poly-(fluorinated alkanesulfonamides) a series of model experiments were carried out: (1) reactions of fluorinated alkanesulfonyl fluorides with amines, (2) reactions of fluorinated alkanesulfonyl chloride with amines and (3) reactions of sodium salts of fluorinated alkanesulfonamides with alkyl iodides of fluorinated alkanesulfonic acid esters. Seventeen new fluorinated alkanesulfonamides were prepared in good yields, namely: RFO(CF2)2SO2NR1R2 (1a-h), R1R2NSO2RFSO2NR1R2 (2a-h) and [Cl (CF2)4O(CF2)2SO2NH(CH2)3]2 (3). Reaction of RFSO2NH2 with equivalent amount of NaOCH3 and methyl iodide was shown to give both the N-mono- and N,N-di-substituted amides. Consequently the N-monosubstituted alkanesulfonamides were chosen as monomers for syntheses of the poly-(fluorinated alkanesulfonamides) and two new polymers were synthesized. The effect of the condition of the polycondensation on M?n of the polymers were discussed and elemental composition, 19F NMR, IR, M?n, Tg, tensile strength, thermal and chemical stabilities of the polymers were measured. Several new perfluoroalkanesulfonyl chlorides CISO2RFSO2Cl (4a-c) and fluorinated alkanesulfonic acid esters (6a-d) were synthesized. However, reaction of CFCl2CF2O(CF2)2SO2F with AlCl3 was found to give Cl3CCF2O(CF2)2SO2F (5) instead of the expected sulfonyl chloride.  相似文献   

6.
For the first time, fluorinated oxathialones, polyfluoroalkylchlorothioformates, chlorocarbonylpolyfluoroalkylsulfenate esters, a chlorocarbonylhexafluoroisopropylidenimino sulfenate, and a 5-tri-fluoromethyl-2-oxo-1,3,4-oxathiazole were synthesized by reacting chlorocarbonylsulfenyl chloride with RfC(O)CH2C(O)R′ (Rf = CF3; R'= CF3, OC2H5), RfO-Li+ (Rf = CF3CH2, (CF3)2C=N-Li+ and CF3C(O)NH2. Perfluorosuccinic acid and mercury(II) trifluoroacetate with ClC(O)SCI gave their respective anhydrides.  相似文献   

7.
Metal Complexes of Biologically Important Ligands, CLVII [1] Halfsandwich Complexes of Isocyanoacetylamino acid esters and of Isocyanoacetyldi‐ and tripeptide esters (?Isocyanopeptides”?) N‐Isocyanoacetyl‐amino acid esters CNCH2C(O) NHCH(R)CO2CH3 (R = CH3, CH(CH3)2, CH2CH(CH3)2, CH2C6H5) and N‐isocyanoacetyl‐di‐ and tripeptide esters CNCH2C(O)NHCH(R1)C(O)NHCH(R2)CO2C2H5 and CNCH2C(O)NHCH(R1)C(O)NHCH (R2)C(O)NHCH(R3)CO2CH3 (R1 = R2 = R3 = CH2C6H5, R2 = H, CH2C6H5) are available by condensation of potassium isocyanoacetate with amino acid esters or peptide esters. These isocyanides form with chloro‐bridged complexes [(arene)M(Cl)(μ‐Cl)]2 (arene = Cp*, p‐cymene, M = Ir, Rh, Ru) in the presence of Ag[BF4] or Ag[CF3SO3] the cationic halfsandwich complexes [(arene)M(isocyanide)3]+X? (X = BF4, CF3SO3).  相似文献   

8.
The pseudohalide CF3SO2NCO has been synthesized by means of a new reaction involving trifluoromethanesulphonamide and chlorosulphonylisocyanate. This method may be used for preparing other perfluorinated alkanesulphonyl-, arenesulphonyl- and alkanecarbonyl-amides. Reactions of CF3SO2NCO with alcohols, thiols, phenols and amines lead to the corresponding carbonic acid esters, thio-esters, phenyl esters and ureas. Reactions with carbonic acids, aldehydes and dimethylsulphoxide gave CO2 and the corresponding acid amides, azomethines and imino-dimethylsulphurane. Under pressure at 160°C, CF3SO2NCO reacts with phosphorus pentasulphide to give the previously unknown compound CF3SO2NCS; with phosphorus pentachloride under the same conditions, CF3SO2NCCl2 is formed.  相似文献   

9.
A series of perfluoropolyether bis‐carboxylic esters was synthesized and their hydrolytic stability investigated. Their formula is ROOCCF2O(CF2CF2O)p(CF2O)qCF2COOR, where p/q = 1.07 and p + q = 2.94. The alkyl group, R, varied both in terms of steric hindrance and electron‐withdrawing ability. Kinetic and thermodynamic data were obtained under homogeneous conditions and compared to a fully hydrogenated ester having a closely related structure CH3(CH2)3OOCCH2O(CH2CH2O)nCH2COO(CH2)3CH3, where n? = 10.6. Neutral ester hydrolysis (NEH) conditions were selected with methyl ethyl ketone as a solvent and a 3–4:1 water/ester ratio. The course of the reaction was monitored by 19F NMR or 1H NMR (when R = CH3CH2? ). Results indicated that the hydrolysis of fluorinated esters, with alkyl aliphatic substituents, is governed by steric hindrance of the substituents. Two distinctive kinetic regimes were observed. The first one, at low conversion, was characterized by lower kinetic constants and related to true NEH conditions. The second regime appeared at higher conversion when acidic autocatalysis dictated the reaction behavior. This is the only observed mechanism when esters more sensitive to the hydrolysis are considered. In these cases, polar factors prevail over steric considerations. Finally, all fluorinated esters of the class (I) showed a much higher reactivity than the hydrogenated ester whose hydrolysis took place only in the presence of a strong acidic catalyst. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 40: 4266–4280, 2002  相似文献   

10.
Organozinc compounds, obtained from dibromomalonic acid dialkyl esters and zinc, react with 2-arylmethylmalonic acid dinitriles and 3-aryl-2-cyanopropenoic acid methyl esters forming 3-aryl-2,2-dicyanocyclopropane-1,1-dicarboxylic acid dialkyl esters and 3-aryl-2-cyanocyclopropane-1,1,2-tricarboxylic acid trimethyl esters. 1H and 13C NMR, 3 J CH constants are considered.  相似文献   

11.
Direct spectroscopic evidence for radiation-induced crosslinking of poly(tetrafluoroethylene) (PTFE) is presented for all x-ray and electron dose levels above which it is possible to distinguish between deliberately introduced radiation damage and the x-ray damage inherent in obtaining an x-ray photoelectron spectrum (XPS). The C (1s) spectrum obtained after irradiation with 2 keV electrons for all doses greater than 1 μA-min/cm2 consists of a four-peak spectrum identical to that previously obtained for plasma-polymerized tetrafluoroethylene and assigned to carbon atoms with variable numbers of bound F atoms (CF3, CF2, CF1, and CF0). X-ray irradiated PTFE can be fitted with the same four-peak spectrum. At or below an electron dose level of 1 μA-min/cm2, the radiation damage is comparable to that produced by the x-ray dose necessary to obtain an XPS spectrum. The CF1 and CF0 components increase with increasing electron dose, and at high electron doses dominate the spectrum. With increasing dose the CF3 component approaches a constant value while both the CF2 component and the total F : C ratio decreases. These four components are those expected to result from radiation-induced crosslinking reactions of the polymer and are consistent with previous suggestions that crosslinking is the basis of radiation patterned adhesion to PTFE. © 1993 John Wiley & Sons, Inc.  相似文献   

12.
The effects of the solvent and the ligand chirality on the regioselectivity of oxidative esterification of propylene and cyclohexene by PdII carboxylates were studied using achiral (MeCO2 , Me2CHCH2CO2 ), racemic ((±)-CF3CF2CF2OC*F(CF3)CO2 ), and chiral ((S)−(+)−MeC*H(Et)CO2 , (+)−CF3CF2CF2OC*F(CF3)CO2 ) carboxylate ligands. The oxidation of alkenes in aprotic media (CHCl3, CH2Cl2, CO2, THF) affords mainly allylic esters (in the case of cyclohexene also homoallylic esters) and the oxidative esterification at the vinylic position is absent. In weakly solvating media (CHCl3, CH2Cl2) the regioselectivity of cyclohexene oxidation (the allyl to homoallyl ratio) increases substantially on going from achiral or racemic acido ligands to chiral acido ligands. In a more donor medium (THF) the ligand chirality effect almost vanishes. The effects of the ligand chirality and the nature of the solvent on the mechanism of alkene oxidation by PdII complexes are discussed. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 9, pp. 1695–1703, September, 1999.  相似文献   

13.
Various aryl-, heteroaryl-, and alkyl mercaptanes (RSH, 1ar) were treated with a slight excess of NaH suspended in DMF to make the appropriate sodium thiolates (RSNa), which then reacted with 1.3 equivalent of CF3I at room temperature for overnight to afford the appropriate trifluoromethyl sulfides (CF3SR, 2) in fair to good yields. The radical chain alkylation reaction was effective without the use of UV irradiation with all but three substrates (thiosalicylic acid, 1k; 2-mercaptobenzimidazole, 1q; and 3-mercaptopropionic acid, 1r).Steam-distillation was found as an effective and easy to upscale means for the isolation of these volatile and water immiscible sulfides. The CF3I reagent gas was conveniently weighed and delivered to the reaction mixture by the balloon technique or as a preliminary made stock solution in DMF or DMSO. The sulfides 2 obtained here were assayed by GC and characterized by 1H, 13C, 19F NMR and MS spectroscopy.  相似文献   

14.
Derivatization yields of different esters (methyl, ethyl, n-propyl, isopropyl, n-butyl and isobutyl) of citric, malic and isocitric acids with various acylating agents, such as acetic anhydride, propionic anhydride, trifluoroacetic anhydride and heptafluorobutyric anhydride, were investigated. The formation of acylated malic and isocitric acid esters is rapid and quantitative whereas the acylation of citric acid esters was slow and partial in most instances. It was found that selective analytical conditions can be achieved with the O-heptafluorobutyryl n-butyl esters. The analytical applicability of the separation and determination of the malic, isocitric and citric acid contents of model solutions and of pressed lemon juice is discussed. Concentrations of 1.2 × 10?3 ?9.0 × 10?3g per 100 g of isocitric acid and 4.2 × 10?3 ?3.5 × 10?2 g per 100 g of malic acid in the presence of 9.3 × 10?1 g per 100 g of citric acid were measured with relative standard deviations of 7.5 and 4.2%, respectively.  相似文献   

15.
To develop the radical polyaddition of bisperfluoroisopropenyl esters to preparation of polymers bearing higher fluorine content, the polyaddition reactivity of bis(α-trifluoromethyl-β,β-difluorovinyl) 2,3,5,6-tetrafluoroterephthalate [CF2C(CF3)OCOC6F4COOC(CF3)CF2] (TFT) with 1,4-dioxane (DOX) and diethoxydimethylsilane (DEOMS) were described. The results of the model reactions of 2-pentafluorobenzoxypentafluoropropene [CF2C(CF3)OCOC6F5] (PFBP) with THF, DOX and DEOMS showed that the reactions took place almost quantitatively and the main products were mono-addition compound for THF and di-addition compounds for DOX and DEOMS, respectively. The polyaddition of TFT with DOX or DEOMS yielded corresponding polymers of about 1×104 as a molecular weight bearing unimodal molecular weight distribution by the initiation of peroxides such as benzoyl peroxide and di-tert-butyl peroxide. TFT showed the slightly higher reactivity compared to that of non-fluorinated analogue, bis(α-trifluoromethyl-β,β-difluorovinyl) terephthalate (BFP), by the results of ternary polyaddition of TFT/BFP/DOX system. Polymers bearing TFT moiety showed the higher thermostability and contact angle.  相似文献   

16.
The synthesis of a variety of new halo-F-methylphosphonates has been achieved by a Michaelis-Arbuzov type reaction between a halo-F-methane and a trialkyl phosphite. This synthesis has proved to be of wide scope and utility for the high yield preparation of a number of heretofore unknown compounds. The 1H, 19F, 13C and 31P NMR spectroscopic properties are reported in detail. The mechanism for the formation of bromodifluoromethylphosphonates has been shown to proceed through the intermediacy of difluorocarbene:CF2. The phosphonate products have been shown to react with a wide variety of reagents. Fluoride and alkoxide ions react by attack at phosphorus with cleavage of the carbon-phosphorus bond and formation of [:CF2] from the bromodifluoromethylphosphonates and the CFBr2 anion from the dibromofluoromethylphosphonates. Iodide ion and tertiary phosphines react by attack at the ester carbon to give stable phosphonate salts. Hydrolysis of the phosphonate esters with 50% aqueous HCl gives the expected phosphonic acids. Trimethylsilyl bromide attacks phosphoryl oxygen to afford the bis(trimethylsilyl) esters.  相似文献   

17.
Abstract

We have previously established that the α-fluorination of alkanephosphonates provides analogues of phosphate esters which have improved ‘isopolarity’ relative to simple alkanephosphonates.1 This property is manifest, inter alia, in enhanced acidity and in the upfield shift for the 31P n.m.r. resonance. Indeed, for a range of halomethanephosphonic acids we have found the relationship “δP=9.61 (pKa2 - 4.59) ppm” gives an excellent correlation between these parameters. In this context, the properties of CF2CIPO(OR)2 species, derived from the Michaelis-Becker reaction of dialkyl phosphonates with Freon 22, CF2Cl2, will be described.  相似文献   

18.
The rate constant for the reaction of the hydroxyl radical with 1,1,1,3,3-pentafluorobutane (HFC-365mfc) has been determined over the temperature range 278–323K using a relative rate technique. The results provide a value of k(OH+CF3CH2CF2CH3)=2.0×10−12exp(−1750±400/T) cm3 molecule−1 s−1 based on k(OH+CH3CCl3)=1.8×10−12 exp (−1550±150/T) cm3 molecule−1 s−1 for the rate constant of the reference reaction. Assuming the major atmospheric removal process is via reaction with OH in the troposphere, the rate constant data from this work gives an estimate of 10.8 years for the tropospheric lifetime of HFC-365mfc. The overall atmospheric lifetime obtained by taking into account a minor contribution from degradation in the stratosphere, is estimated to be 10.2 years. The rate constant for the reaction of Cl atoms with 1,1,1,3,3-pentafluorobutane was also determined at 298±2 K using the relative rate method, k(Cl+CF3CH2CF2CH3)=(1.1±0.3)×10−15 cm3 molecule−1 s−1. The chlorine initiated photooxidation of CF3CH2CF2CH3 was investigated from 273–330 K and as a function of O2 pressure at 1 atmosphere total pressure using Fourier transform infrared spectroscopy. Under all conditions the major carbon-containing products were CF2O and CO2, with smaller amounts of CF3O3CF3. In order to ascertain the relative importance of hydrogen abstraction from the (SINGLE BOND)CH2(SINGLE BOND) and (SINGLE BOND)CH3 groups in CF3CH2CF2CH3, rate constants for the reaction of OH radicals and Cl atoms with the structurally similar compounds CF3CH2CCl2F and CF3CH2CF3 were also determined at 298 K k(OH+CF3CH2CCl2F)=(8±3)×10−16 cm3 molecule−1 s−1; k(OH+CF3CH2CF3)=(3.5±1.5)×10−16 cm3 molecule−1 s−1; k(Cl+CF3CH2CCl2F)=(3.5±1.5)×10−17 cm3 molecule−1 s−1]; k(Cl+CF3CH2CF3)<1×10−17 cm3 molecule−1 s−1. The results indicate that the most probable site for H-atom abstraction from CF3CH2CF2CH3 is the methyl group and that the formation of carbonyl compounds containing more than a single carbon atom will be negligible under atmospheric conditions, carbonyl difluoride and carbon dioxide being the main degradation products. Finally, accurate infrared absorption cross-sections have been measured for CF3CH2CF2CH3, and jointly used with the calculated overall atmospheric lifetime of 10.2 years, in the NCAR chemical-radiative model, to determine the radiative forcing of climate by this CFC alternative. The steady-state Halocarbon Global Warming Potential, relative to CFC-11, is 0.17. The Global Warming Potentials relative to CO2 are found to be 2210, 790, and 250, for integration time-horizons of 20, 100, and 500 years, respectively. © 1997 John Wiley & Sons, Inc.  相似文献   

19.
The kinetics of the self-reactions of HO2, CF3CFHO2, and CF3O2 radicals and the cross reactions of HO2 with FO2, HO2 with CF3CFHO2, and HO2 with CF3O2 radicals, were studied by pulse radiolysis combined with time resolved UV absorption spectroscopy at 295 K. The rate constants for these reactions were obtained by computer simulation of absorption transients monitored at 220, 230, and 240 nm. The following rate constants were obtained at 295 K and 1000 mbar total pressure of SF6 (unit: 10−12 cm3 molecule−1 s−1): k(HO2+HO2)=3.5±1.0, k(CF3CFHO2+CF3CFHO2)=3.5±0.8, k(CF3O2+CF3O2)=2.25±0.30, k(HO2+FO2)=9±4, k(CF3CFHO2+HO2)=5.0±1.5, and k(CF3O2+HO2)=4.0±2.0. In addition, the decomposition rate of CF3CFHO radicals was estimated to be (0.2–2)×103 s−1 in 1000 mbar of SF6. Results are discussed in the context of the atmospheric chemistry of hydrofluorocarbons. © 1997 John Wiley & Sons, Inc.  相似文献   

20.
The thermal gas-phase reaction of CF3OF with CCl2CCl2 has been studied between 313.8 and 343.8 K. The initial pressure of CF3OF was varied between 10.8 and 77.5 torr and that of CCl2CCl2 between 3.7 and 26.8 torr. CF3OF was always present in excess, varying the initial ratio of CF3OF to that of CCl2CCl2 from 1.3 to 10. Three products were formed: CF3OCCl2CCl2F, CCl2FCCl2F, and CF3O(CCl2CCl2) 2OCF3. The yields of CF3OCCl2CCl2F were 98–99.5%, based on the sum of the products. The reaction was a homogeneous chain reaction not affected by the total pressure. In presence of O2 the oxidation of CCl2CCl2 to CCl3C(O)Cl and COCl2 occurred. The proposed basic reaction steps are: generation of the radicals CF3O˙ and CCl2FCCl2˙ (κ1) in a biomolecular process between CF3OF and CCl2CCl2, formation of the radical CF3OCCl2CCl2˙ by addition of CF3O˙ to CCl2CCl2, chain generation of CF3O˙ by abstraction of fluorine atom from CF3OF by CF3OCCl2CCl2˙ (κ4), and chain termination by recombination of the radicals CF3OCCl2CCl2˙. The expressions obtained for the constants κ1 and κ4 are κ1 = 3.16 ± 0.6 × 107 exp(−15.2 ± 1.7 Kcal mol−1/RT) dm3 mol−1 s−1, κ4 = 3.7 ± 0.5 × 109 exp(−6.0 ± 1.1 Kcal mol−1/RT) dm3 mol−1 s−1. © 1996 John Wiley & Sons, Inc.  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