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.
Merck KGaA, Germany recently tested a new electrolyte salt LiPF3(CF3CF2)3 (lithium fluoroalkyl phosphate (LiFAP)) for lithium ion power packs and suggested that it can be replaced with commercially used LiPF6. LiFAP, for the first of its kind, was incorporated into polyvinylidenefluoride-co-hexafluoropropylene (PVdF-HFP) matrix with ethylene carbonate and diethyl carbonate mixtures as plasticizing agents and SiO2 nanoparticles as filler. Nanocomposite polymer electrolyte (NCPE) membranes were prepared by solvent casting technique. All NCPE membranes were subjected to a.c. impedance, fluorescence and morphological studies. NCPE membranes containing 2.5 wt% of SiO2 exhibited enhanced conductivity of 1.13 mS cm−1 at ambient temperature. Molecular motion in the polymeric media was supported by fluorescence studies. The percentage of crystallinity and activation energy has also been calculated.  相似文献   

2.
The rate constants for the reactions of OH radicals with CH3OCF2CF3, CH3OCF2CF2CF3, and CH3OCF(CF3)2 have been measured over the temperature range 250–430 K. Kinetic measurements have been carried out using the flash photolysis, laser photolysis, and discharge flow methods combined respectively with the laser induced fluorescence technique. The influence of impurities in the samples was investigated by using gas‐chromatography. The following Arrhenius expressions were determined: k(CH3OCF2CF3) = (1.90) × 10−12 exp[−(1510 ± 120)/T], k(CH3OCF2CF2CF3) = (2.06) × 10−12 exp[−(1540 ± 80)/T], and k(CH3OCF(CF3)2) = (1.94) × 10−12 exp[−(1450 ± 70)/T] cm3 molecule−1 s−1. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 846–853, 1999  相似文献   

3.
Li insertion–deinsertion into composite graphite electrodes, comprising synthetic graphite flakes (6 μm average size), polyvinylidene difluoride binder (PVdF), and copper current collectors, in commonly used alkyl carbonate solutions were studied by in situ atomic force microscopy (AFM). In this study, we were able to probe by in situ AFM the behavior of practical, composite graphite electrodes in ethylene carbonate–dimethyl carbonate (EC–DMC) solutions containing salts such as LiAsF6 and LiPF6 during entire lithiation–delithiation cycles. These in situ micro/nanomorphological studies could probe surface film formation on the graphite particles, as well as periodic volume changes in the graphite flakes during Li insertion–deinsertion cycles. These cyclic volume changes can explain the capacity fading of graphite electrodes upon prolonged cycling, in Li-ion batteries. While the overall morphology of these electrodes remains steady upon cycling in the appropriate solutions (in which the Li–C electrodes are efficiently passivated), there is a continuous problem in the extent of accommodation of the small volume changes in the graphite particles upon lithiation–delithiation, by the surface films. It is suggested that graphite electrodes fail during prolonged cycling due to small scale, continuous reactions of the active mass with solution species, which gradually increase their impedance and decrease the content of the lithium stored in the electrodes.  相似文献   

4.
The kinetics of the reactions O(3P) + CF2CCl2 and O(3P) + CF3CFCF2 were studied at room temperature in a discharge flow tube system. The overall rate constants based on the measured afterglow reactions were (3.10 ± 0.40) × 10−13 and (3.00 ± 0.60) × 10−14 cm3 molecule−1 s−1, respectively. The experiments were carried out under pseudo‐first‐order conditions with [O(3P)]0 ≪ [alkene]0. These results are compared with previous relative measurements using different experimental techniques. The effect of substituent atoms or groups on the overall rate constants is analyzed in comparison with other alkenes in the literature. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 867–872, 1999  相似文献   

5.
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   

6.
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   

7.
HPLC separation of the products of high-temperature reaction of a sublimed mixture of C60–C70 (10: 1) with CF3I in a sealed ampoule allowed isolation and determination of molecular structures (X-ray crystallography and 19F NMR) of two new isomers of C60(CF3)12 and one isomer of C60(CF3)14. These isomers are characterized by low relative formation energies, which suggests that the trifluoromethylation process is basically under the thermodynamic control.  相似文献   

8.
Synthesis of Carboxylate Substituted Rhenium Gold Metallatetrahedranes Re2(AuPPh3)2(μ-PCy2)(CO)71-OC(R)O) (R = H, Me, CF3, Ph, 3,4-(OMe)2C6H3) The reaction of the in situ prepared salt Li[Re2(μ-H)(μ-PCy2)(CO)7(ax-C(Ph)O)] ( 2 ) with 1,5 equivalents of monocarboxylic acid RCOOH (R = H ( 4 a ), Me ( 4 b ), CF3 ( 4 c ), Ph ( 4 d ), 3,4-(OMe)2C6H3 ( 4 e ) in tetrahydrofruan (THF) solution at 60 °C gives within 4 h under release of benzaldehyde (PhCHO) the η1-carboxylate substituted dirhenium salt Li[Re2(μ-H)(μ-PCy2)(CO)71-OC(R)O)] (R = H ( 4 a ), Me ( 4 b ), CF3 ( 4 c ), Ph ( 4 d ), 3,4-(OMe)2C6H3 ( 4 e )) in almost quantitative yield. The lower the pKa value of the respective carboxylic acid the faster the reaction proceeds. It was only in the case of CF3COOH possible to prove the formation of the hydroxycarbene complex Re2(μ-H)(μ-PCy2)(CO)7(=C(Ph)OH) ( 5 ) prior to elimination of PhCHO. The new compounds 4 a–4 e were only characterized by 31P NMR and ν(CO) IR spectroscopy as they are only stable in solution. They are converted with two equivalents of BF4AuPPh3 at 0 °C in a so-called cluster expansion reaction into the heterometallic metallatetrahedrane complexes Re2(AuPPh3)2(μ-PCy2)(CO)71-OC(R)O) (R = H ( 7 a ), Me ( 7 b ), CF3 ( 7 c ), Ph ( 7 d ), 3,4-(OMe)2C6H3 ( 7 e )) (yield 47–71% ). The expected precursor complexes of 7 a–7 e Li[Re2(AuPPh3)(μ-PCy2)(CO)71-OC(R)O] ( 8 ) were not detected by NMR and IR spectroscopy in the course of the reaction. Their existence was retrosynthetically proved by the reaction of 7 b with an excess of the chelating base TBD (1,5,7-Triazabicyclo[4.4.0]dec-5-en) forming [(TBD)xAuPPh3][Re2(AuPPh3)(μ-PCy2)(CO)71-OC(Me)O] ( 8 b ) in solution. The η1-bound carboxylate ligand in 7 a–7 e can photochemically be converted into a μ-bound ligand in Re2(AuPPh3)2(μ-PCy2)(μ-OC(R)O)(CO)6 (R = H ( 9 a ), Me ( 9 b ), CF3 ( 9 c ), Ph ( 9 d ), 3.4-(MeO)2C6H3 ( 9 e )) under release of one equivalent CO. All isolated cluster complexes were characterized and identified by the following analytical methods: elementary analysis, NMR (1H, 31P) spectroscopy, ν(CO) IR spectroscopy and in the case of 7 d and 9 b by X-ray structure analysis.  相似文献   

9.
Exposure of the tetrameric, heterocubane‐like perfluorinated lithium alkoxide [Li{OC(CF3)3}]4 to humid air gaverise to the hydrolysis products [{(CF3)3CO}Li(H2O)2μ‐(H2O)‐Li(H2O)2{OC(CF3)3}], [{(CF3)3CO}Li(H2O)2μ‐(H2O)‐Li‐(H2O)3]+[OC(CF3)3] and [Li(H2O)4]+[OC(CF3)3] because of stepwise addition of water molecules in a gas‐solid reaction without solvent. All compounds were studied by X‐ray crystallography and their solid‐state structures are strongly influenced by hydrogen bonding and fluorophilic interactions.  相似文献   

10.
The processes of vibrational relaxation and unimolecular dissociation of the perfluoromethyl halides CF3Cl, CF3Br, and CF3I have been studied in the shock tube with the laser-schlieren technique. Vibrational relaxation was resolved in pure CF3Cl and CF3Br (400–484 K and 400–500 K, respectively), and in the mixtures; 2% CF3Cl/Kr (500–1000 K), 10% CF3Cl/Kr (440–670 K), 4% CF3Br/Kr (450–850 K), and 2% CF3I/Kr (620–860 K). Relaxation in the pure gases is extremely rapid, but shows a well-resolved, accurately exponential decay which provides very precise relaxation times in close agreement with ultrasonic results. Relaxation times as short as 0.1 μs-atm can be resolved, showing the method has a resolution within a factor 2–3 of the best ultrasonic methods. Relaxation dilute in rare gas shows a complex double exponential behavior consistent with a two-stage series process. Rates of CF3(SINGLEBOND)X fission in these mixtures were measured over 1800–3000 K, P<0.55 atm, for CF3Cl; 1600–2500 K, P<0.55 atm, in CF3Br; and 1260–2100 K, P<0.34 atm, in CF3I. Rates for dissociation were derived from a full profile modeling using a secondary mechanism of six CF3 reactions. RRKM analysis showed all dissociations to lie near the low pressure limit. Using literature barriers, these rates are best fit with (ΔE)all=−270 cm−1 for CF3Cl, 〈ΔEdown=0.3 T for CF3Br, and 〈ΔEdown=800 cm−1 for CF3F. All these transfers are on the large side, similar to those found in other halogenated methanes. © 1997 John Wiley & Sons, Inc.  相似文献   

11.
Rate constants have been determined for the reactions of Cl atoms with the halogenated ethers CF3CH2OCHF2, CF3CHClOCHF2, and CF3CH2OCClF2 using a relative‐rate technique. Chlorine atoms were generated by continuous photolysis of Cl2 in a mixture containing the ether and CD4. Changes in the concentrations of these two species were measured via changes in their infrared absorption spectra observed with a Fourier transform infrared (FTIR) spectrometer. Relative‐rate constants were converted to absolute values using the previously measured rate constants for the reaction, Cl + CD4 → DCl + CD3. Experiments were carried out at 295, 323, and 363 K, yielding the following Arrhenius expressions for the rate constants within this range of temperature:Cl + CF3CH2OCHF2: k = (5.15 ± 0.7) × 10−12 exp(−1830 ± 410 K/T) cm3 molecule−1 s−1 Cl + CF3CHClOCHF2: k = (1.6 ± 0.2) × 10−11 exp(−2450 ± 250 K/T) cm3 molecule−1 s−1 Cl + CF3CH2OCClF2: k = (9.6 ± 0.4) × 10−12 exp(−2390 ± 190 K/T) cm3 molecule−1 s−1 The results are compared with those obtained previously for the reactions of Cl atoms with other halogenated methyl ethyl ethers. © 2001 John Wiley & Sons, Inc. Int J Chem Kinet 33: 165–172, 2001  相似文献   

12.
The i.r. gas and Raman liquid spectra of CF3Si(CH3)3 and CF3Si(CD3)3 are reported and assigned for C3vsymmetry. Force constants have been calculated by a combined analysis of both isotopomers yielding ƒ (SiCF3) 2.63, ƒ (SiCH3) 3.07 and ƒ (CF) 5.70 N cm−1. The apparent weakness of the SiCF3 bond confirms the results obtained on other CF3 silanes and is discussed with respect to related molecules.  相似文献   

13.
《Mendeleev Communications》2022,32(2):212-214
The compound Na2Ti(CF3COO)6(CF3COOH)2 was synthesized as colorless crystals, extremely unstable in air, by the reaction of TiCl4 with trifluoroacetic acid and sodium trifluoroacetate. Crystallographic studies have shown that this sodium trifluoroacetatotitanate is the first example of a tetravalent titanium carboxylate [Ti(RCOO)6]2– containing titanium in an octahedral environment of oxygen atoms of carboxylate groups. Thermal decomposition of Na2Ti(CF3COO)6(CF3COOH)2 in an argon atmosphere results in the complex fluoride Na2TiF6.  相似文献   

14.
Syntheses and Characterizations of the First Tris and Tetrakis(trifluoromethyl) Palladates(II) and Platinates(II), [M(CF3)3(PPh3)] and [M(CF3)4]2— (M = Pd, Pt) Tris(trifluoromethyl)(triphenylphosphino)palladate(II) and platinate(II), [M(CF3)3PPh3], and the tetrakis(trifluoromethyl)metallates, [M(CF3)4]2— (M = Pd, Pt), are prepared from the reactions of [MCl2(PPh3)2] and Me3SiCF3 / [Me4N]F or [I(CF3)2] salts in good yields. [Me4N][M(CF3)3(PPh3)] crystallize isotypically in the orthorhombic space group Pnma (no. 62) with Z = 4. The NMR spectra of the new compounds are described.  相似文献   

15.
In the system LiSO3CF3/RbSO3CF3 four different quasi‐ternary phases occur: Li0.7Rb0.3SO3CF3, Li0.55Rb0.45SO3CF3, LiRb2(SO3CF3)3, and Li0.2Rb0.8SO3CF3. These have been identified, and characterized by means of X‐ray powder diffractometry and DSC. LiSO3CF3 is trimorphic, LiRb2(SO3CF3)3 is dimorphic and RbSO3CF3 exists in four different modifications. The cation dynamics has been studied using 7Li‐NMR line shape analysis and 7Li‐spin lattice relaxation (T1) measurements. The pure and mixed trifluoromethylsulfonates in the system LiSO3CF3/RbSO3CF3 are solid electrolytes. Their ionic conductivities below 475 K increase with the rubidium content. Above this temperature, the conductivity of β‐LiRb2(SO3CF3)3 exceeds the one of δ‐RbSO3CF3.  相似文献   

16.
On Sn[OCH(CF3)2]2 and Sn(OCH2CF3)2 (n = 1, 2) The sulfoxylates S[OCH(R)CF3]2, 1 and 2 and the disulfides S2[OCH(R)CF3]2, 5 and 6 (R = CF3, H) are obtained by reacting SCl2 or S2Cl2, respectively, and the lithium alcoxides LiOCH(R)CF3. Chlorine and compound 2 give ClS(O)OCH2F3 and CF3CH2Cl, whereas the sulfur-sulfur bound is cleaved in 5 and 6 furnishing SCI2, 1 and 2 , respectively. The 19F n.m.r. spectrum of 5 and the 1H n.m.r. spectrum of 6 are interpreted in terms of hindered rotation about the sulfur-sulfur axis.  相似文献   

17.
(CF3)2PAsH2 and (CF3)2AsAsH2 (CF3)2PAsH2 is obtained in yields between 30 and 60% according to eq. (1) (CF3)2AsAsH2 is formed by the analogous reaction with (CF3)2AsI, but is not sufficiently stable to be isolated. Both compounds are decomposed according to eq. (2) (CF3)2PAsH2 can be studied in solution below ?40°C; it is characterized by molar mass determination and by its n.m.r. spectra (1H, 19F, 31P). Reactions with polar [HBr, (CH3)2AsH, (CH3)2PN(CH3)2] and nonpolar [Br2, As2(CH3)4] reagents proceed by cleavage of the P? As bond.  相似文献   

18.
The title compounds, poly­[[[bis(2‐methoxy­ethyl) ether]­lithium(I)]‐di‐μ3‐tri­fluoro­methanesulfonato‐lithium(I)], [Li2(CF3SO3)2(C6H14O3)]n, and poly­[[[bis(2‐methoxy­ethyl) ether]­lithium(I)]‐di‐μ3‐tri­fluoro­acetato‐dilithium(I)‐μ3‐tri­fluoro­acetato], [Li3(C2F3O2)3(C6H14O3)]n, consist of one‐dimensional polymer chains. Both structures contain five‐coordinate Li+ cations coordinated by a tridentate diglyme [bis(2‐methoxy­ethyl) ether] mol­ecule and two O atoms, each from separate anions. In both structures, the [Li(diglyme)X2]? (X is CF3SO3 or CF3CO2) fragments are further connected by other Li+ cations and anions, creating one‐dimensional chains. These connecting Li+ cations are coordinated by four separate anions in both compounds. The CF3SO3? and CF3CO2? anions, however, adopt different forms of cation coordination, resulting in differences in the connectivity of the structures and solvate stoichiometries.  相似文献   

19.
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   

20.
The synthesis, IR spectrum, and first‐principles characterization of CF3CH(ONO)CF3 as well as its use as an OH radical source in kinetic and mechanistic studies are reported. CF3CH(ONO)CF3 exists in two conformers corresponding to rotation about the RCO? NO bond. The more prevalent trans conformer accounts for the prominent IR absorption features at frequencies (cm?1) of 1766 (N?O stretch), 1302, 1210, and 1119 (C? F stretches), and 761 (O? N? O bend); the cis conformer contributes a number of distinct weaker features. CF3CH(ONO)CF3 was readily photolyzed using fluorescent blacklamps to generate CF3C(O)CF3 and, by implication, OH radicals in 100% yield. CF3CH(ONO)CF3 photolysis is a convenient source of OH radicals in the studies of the yields of CO, CO2, HCHO, and HC(O)OH products which can be difficult to measure using more conventional OH radical sources (e.g., CH3ONO photolysis). CF3CH(ONO)CF3 photolysis was used to measure k(OH + C2H4)/k(OH + C3H6) = 0.29 ± 0.01 and to establish upper limits of 16 and 6% for the molar yields of CO and HC(O)OH from the reaction of OH radicals with benzene in 700 Torr of air at 296 K. © 2003 Wiley Periodicals, Inc. Int J Chem Kinet 35: 159–165, 2003  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