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 75 毫秒
1.
The kinetics of the decomposition of benzotrifluoride was studied from 720°c to 859°c in a flow system with and without carrier gas. Consideration of the product distribution made possible the study of the decomposition into CF3 and C6H5 radicals, which appeared to be truly homogeneous in character. The first-order rate constant of the C? C bond fission, log k (sec?1) = (17.9 ± 0.5) (99.7 ± 2.5)/θ, did not change with change of initial concentration, pressure of the carrier gas, or contact time. The Arrhenius parameters have been related to the appropriate thermodynamic data. Assumption of 0 kcal/mole for the activation energy of the reverse combination reaction yielded DH(C6H5? CF3) = 103.6 ± 2.5 kcal/mole and ΔH(C6H5) = 77.1 ± 3.0 kcal/mole. Applicability of the simple first-order formula to calculation of the rate constant has been also dealt with.  相似文献   

2.
The high-pressure absolute rate constants for the decomposition of nitrosobenzene and pentafluoronitrosobenzene were determined using the very-low-pressure pyrolysis (VLPP) technique. Bond dissociation energies of DH0(C6H5? NO) = 51.5 ± 1 kcal/mole and DH0 (C6F5? NO) = 50.5 ± 1 kcal/mole could be deduced if the radical combination rate constant is set at log kr(M?1·sec?1) = 10.0 ± 0.5 for both systems and the activation energy for combination is taken as 0 kcal/mole at 298°K. δHf0(C6H5NO), δHf0(C6F5NO), and δHf0(C6F5) could be estimated from our kinetic data and group additivity. The values are 48.1 ± 1, –160 ± 2, and – 130.9 ± 2 kcal/mole, respectively. C–X bond dissociation energies of several perfluorinated phenyl compounds, DH0(C6F5–X), were obtained from the reported values of δHf0(C6F5X) and our estimated δHf0(C6F5) [X = H, CH3, NO, Cl, F, CF3, I, and OH].  相似文献   

3.
The Arrhenius parameters for the title reaction have been measured in a very-low-pressure pyrolysis apparatus in the temperature range 644–722 K and are given by log k2 (M?1 · sec?1) = 9.68 - 2.12/θ, where θ = 2.303RT in kcal/mol. Together with the published Arrhenius parameters for the reverse reaction from iodination studies, they result in a standard heat of formation of the t-butyl radical of 8.4 kcal/mol, accepting S0(C4H9·) = 72.2 e.u. at 300 K from other kinetic data, and thus confirm the accepted value for ΔHf0(t-C4H9·), at variance with recent investigations which yielded significantly higher values. This value for ΔHf0(t-C4H9·) results in a bond-dissociation energy (BDE) for isobutane of 92.7 kcal/mol.  相似文献   

4.
Using the ‘permutation of indices’ method proposed by Kaplan and Fraenkel, we could formulate the density-matrix equations required to fit the temperature-dependent 13C-NMR spectra observed with the title compounds. For 6Li13CHBr2 ( 1 ) and 6Li13CH2SC6H5 ( 2 ) an exchange mechanism is proposed by which monomers interchange C- and Li-atoms via a non-observed dimeric intermediate; the activation parameters of these intermolecular dynamic processes have been found to be ΔH = 10.2 kcal/mol, ΔS = 13.7 cal/mol·K for 1 and ΔH = 11.1 kcal/mol, ΔS = 20.6 cal/mol·K for 2 ((D8)THF as solvent). In the case of (6Li)butyllithium ( 3 ), the observed low-temperature spectra indicate that dimeric ( 3b ) and tetrameric ( 3a ) species are in dynamic equilibrium interchanging the C3HCH2 groups (and THF molecules) bonded to the 6Li-atoms. The relative concentrations of the dimer and of the tetramer have been determined by peak integration or by line-shape fitting; the ‘pseudo’- equilibrium constant, defined by Keq = [ 3b ]2/[ 3a ], was found to be 2.6·10?2 mol/1 (at ?88°) and corresponds to ΔGR (?88°) = 2 ΔG°f( 3b ) – ΔG°f( 3a ) = 1.34 kcal/mol. The activation parameters of the dynamic process responsible for the exchange were estimated as ΔH = 3.78 kcal/mol and ΔS = ?31.3 cal/mol·K. Tentative interpretation of the thermodynamic and kinetic parameters is given.  相似文献   

5.
Kinetics for reactions of phenoxy radical, C6H5O, with itself and with O3 were examined at 298 K and low pressure (1 Torr) using discharge flow coupled with mass spectrometry (DF/MS). The rate constant for the phenoxy radical self‐reaction was determined to be k1 = (1.49 ± 0.53) × 10−11 cm3 molecule−1 s−1 defined by d[C6H5O]/dt=−2 k1[C6H5O]2. The rate constant for the C6H5O reaction with O3 was measured to be k2 = (2.86 ± 0.35) × 10−13 cm3 molecule−1 s−1, which may be a lower limit value. Because of much higher atmospheric abundance of ozone than that of both NO and phenoxy, the reaction of C6H5O with ozone may represent the principal fate of the phenoxy radical in the atmosphere. Products from reaction of C6H5O + C6H5O, NO, and NO2 were also investigated, and (C6H5O)2 (m/e = 186), C6H5O(NO) (m/e = 123), and C6H5O(NO2) (m/e = 139) adducts were observed as products for the reactions of C6H5O with itself, NO, and NO2, respectively. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 65–72, 1999  相似文献   

6.
Proton decoupled 13C NMR spectra have been measured for the cyclopentadienyl compounds C5H5Si(CH3)nCl3?n(n = 1, 2, 3), C5H5Ge(CH3)3, CH3C5H4Ge(CH3)3, C5H5Sn(CH3)3, σ-C5H5Fe(CO)2-π-C5H5 and C5H5HgCH3. A fast metallotropic rearrangement occurring in the compounds causes the spectra to be temperature dependent for the Si, Ge, Sn and Fe derivatives. For the derivatives of silicon or germanium, the olefinic signals are unsymmetrically broadened by the 1,2-shift at lower migration rates. Line widths of the ring carbon signals have been measured to give an estimate for the activation parameters of the rearrangement in C5H5Ge(CH3)3 (Ea = 10·7 ± 0·9 kcal/mole, ΔG? = 13·4 ± 0·9 kcal/mole) and C5H5Sn(CH3)3 (Ea = 6·8 ± 0·7 kcal/mole, ΔG? = 7·1 ± 0·7 kcal/mole). At room temperature, the spectrum of C5H5HgCH3 displays just one narrow signal responsible for the cyclopentadienyl ligand. The spectrum of CH3C5H4Ge(CH3)3 at –30° demonstrates that two isomers containing methyl in the vinylic position are present, the ratio being ca. 2:1. The 13C spectra of the vinylic isomers have been analysed in the case of C5H5Si(CH3)nCl3?n.  相似文献   

7.
Using an oxenoid model, we investigated dependences of carcinogenic potency of the benzenes C6H5‐X on a nature of substituents X. According to the model, a P450 enzyme breaks a dioxygen molecule and generates the oxens, which readily react with substrates. We suggest that a stability of the intermediate OC6H6‐X with tetrahedrally coordinated C atom relative to the molecule C6H5‐X determines a rate of substrate biotransformation. Using MO LCAO MNDO approach, we calculated the total energies of molecules C6H6‐X and arene oxides OC6H6‐X. A difference ΔEmin of these values determines activation energy of oxidation reaction. The compounds with the low ΔEmin values are noncarcinogenic. Benzene derivatives with high ΔEmin values belong to carcinogenic compounds series. The carcinogenicity of amino‐ and nitro‐substituted benzenes is also determined by N‐oxidation of amino and reduction of the nitro group. As the phenylhydroxylamines XC6H4NHOH and nitrenium ions XC6OH4NH+ are the common metabolites of the nitro‐ and amino‐substituted benzenes and nitrenium ions XC6H4NH+ are the ultimate carcinogens, we use the differences ΔEN = E(XC6H4NH+) ? E(XC6H4NHOH) as the second parameter characterizing the carcinogenic activity of amino‐ and nitro‐substituted benzenes. © 2009 Wiley Periodicals, Inc. Int J Quantum Chem, 2010  相似文献   

8.
Stable N‐heterocyclic carbene analogues of Thiele and Chichibabin hydrocarbons, [(IPr)(C6H4)(IPr)] and [(IPr)(C6H4)2(IPr)] ( 4 and 5 , respectively; IPr=C{N(2,6‐iPr2C6H3)}2CHCH), are reported. In a nickel‐catalyzed double carbenylation of 1,4‐Br2C6H4 and 4,4′‐Br2(C6H4)2 with IPr ( 1 ), [(IPr)(C6H4)(IPr)](Br)2 ( 2 ) and [(IPr)(C6H4)2(IPr)](Br)2 ( 3 ) were generated, which respectively afforded 4 and 5 as crystalline solids upon reduction with KC8. Experimental and computational studies support the semiquinoidal nature of 5 with a small singlet?triplet energy gap ΔES?T of 10.7 kcal mol?1, whereas 4 features more quinoidal character with a rather large ΔES?T of 25.6 kcal mol?1. In view of the low ΔES?T, 4 and 5 may be described as biradicaloids. Moreover, 5 has considerable (41 %) diradical character.  相似文献   

9.
The following reactions: (1) were studied over the temperature ranges 533–687 K, 563–663 K, and 503–613 K for the forward reactions respectively and over 683–763 K, for the back reaction. Arrhenius parameters for chlorine atom transfer were determined relative to the combination of the attacking radicals. The ΔHr°(1) = ?3.95 ± 0.45 kcal mol?1 was calculated and from this value the ΔH∮(C2F5Cl) = ?2.66.3 ± 2.5 kcal mol?1 and D(C2F5-Cl) = 82.0 ± 1.2 kcal mol?1 were obtained. Besides, the ΔHr°(2) was estimated leading to D(CF2ClCF2Cl) = 79.2 ± 5 Kcal mol?1. The bond dissociation energies and the heat of formation are compared with those of the literature. The effect of the halogen substitutents as well as the importance of the polar effects for halogen transfer processes are discussed.  相似文献   

10.
Zusammenfassung Verbindungen des Typs R2NBCIN(Sime 3)2 * (R=C2H5,iC3H7, C6H11, C6H5) bleiben bis zu hohen Temperaturen beständig. Sie reagieren jedoch mit NaN(Sime 3)2 unter Übertragung einer Methylgruppe von einem Silicium-zu einem Boratom zu N,N-Bis(trimethylsilyl)-tetramethylcyclodisildiazan und R2NBmeN(Sime 3) (R=C2H5,iC3H7).
Compounds of the type R2NBClN(Sime 3)2 * were prepared. (R=C2H5,iC3H7, C6H11, C6H5. They do not show any tendency to condense thermally under elimination of trimethylchlorosilane. When these silylaminoboranes were allowed to react with NaN(Sime 3)2, complicated rearrangements were observed. The compounds in which R=C2H5 oriC3H7 rearranged under formation of N,N-bis(trimethylsilyl)-tetramethylcyclodisildiazane and R2NBmeN(Sime)2. In this reaction a migration of a methyl group from silicon to boron occurs.
  相似文献   

11.
Complexes of Titanium — Synthesis, Structure, and Fluxional Behaviour of CpTi{η6‐C5H4=C(p‐Tol)2}Cl (Cp′ = Cp*, Cp) The reaction of Cp′TiCl3 (C′ = Cp* or Cp) with magnesium and 6, 6‐di‐para‐tolylpentafulvene generates good yields of pentafulvene complexes Cp*Ti{η6‐C5H4=C(p‐Tol)2}Cl ( 4 ) and CpTi{η6‐C5H4=C(p‐Tol)2}Cl ( 5 ), respectively. The crystal and molecular structure of 4 have been determined from X‐ray data and exhibits compared to known η6‐pentafulvene complexes an unusual large Ti—C(p‐Tol)2 (Fv)‐distance (2.535(5)Å) evoked by the bulky substituents at the exocyclic carbon. Dynamic 1H‐NMR and spin saturation transfer experiments point out a rotation of the fulvene ligand around the Ti—Ct2 axis (Ct2 = centroid of the fulvene ring carbon atoms) with an activation barrier ΔGC = 60.6 ± 0.5 kJ mol−1 (TC = 314 ± 2 K). For 5 this barrier is significantly larger. Analogous dynamic behaviour is well known for diene complexes, but to our knowledge, it is here first‐time described for a pentafulvene complex.  相似文献   

12.
The kinetics of the bromate ion-iodide ion-L-ascorbic acid clock reaction was investigated as a function of temperature and pressure using stopped-flow techniques. Kinetic results were obtained for the uncatalyzed as well as for the Mo(VI) and V(V) catalyzed reactions. While molybdenum catalyzes the BrO-I? reaction, vanadium catalyzes the direct oxidation of ascorbic acid by bromate ion. The corresponding rate laws and kinetic parameters are as follows. Uncatalyzed reaction: r2 = k2[BrO] [I?][H+]2, k2 = 38.6 ± 2.0 dm9 mol?3 s?1, ΔH? = 41.3 ± 4.2 kJmol?1, ΔS? = ?75.9 ± 11.4 Jmol?1 K?1, ΔV? = ?14.2 ± 2.9 cm3 mol?1. Molybdenum-catalyzed reaction: r2 = k2[BrO] [I?] [H+]2 + kMo[BrO] [I?] [ H+]2[M0(VI)], kMo = (2.9 ± 0.3)106 dm12 mol?4 s?1, ΔH? = 27.2 ± 2.5 kJmol?1, ΔS? = ?30.1 ± 4.5 Jmol?1K?1, ΔV? = 14.2 ± 2.1 cm3 mol?1. Vanadium-catalyzed reaction: r1 = kV[BrO] [V(V)], kV = 9.1 ± 0.6 dm3 mol?1 s?1, ΔH? = 61.4 ± 5.4 kJmol?1, ΔS? = ?20.7 ± 3.1 Jmol?1K?1, ΔV? = 5.2 ± 1.5 cm3 mol?1. On the basis of the results, mechanistic details of the BrO-I? reaction and the catalytic oxidation of ascorbic acid by BrO are elaborated. © 1995 John Wiley & Sons, Inc.  相似文献   

13.
The reaction of benzoyl ion with ammonia in multiple-collision conditions in the second quadrupole assembly of a triple quadrupole mass spectrometer at (laboratory) ion kinetic energies from 0 to 20 eV produced the even-electron ions [C6H5]+, [C6H5NH3·(NH3)m]+ (m = 0, 1) and [C6H5CONH3·(NH3)n]+ (n = 0, 1, 2, 3) and the odd-electron ions [C6H4NH3·(NH3)p]+· (p = 0, 1). Thermochemical information could not be obtained under multiple-collision conditions: both exothermic and endothermic reactions were observed, with no translational-energy onset measurable for the endothermic processes, nor decrease in the yield of the exothermic processes at high energies. The behaviour of cluster-ion intensities as pressure varied was qualitatively as expected. There are pressure and energy regions where spectra change little; if this feature were to be general, it would point to some utility for these conditions in qualitative analysis.  相似文献   

14.
The product from reaction of lanthanum chloride heptahydrate with salicylic acid and thioproline, [La(Hsal)2•(tch)]•2H2O, was synthesized and characterized by IR, elemental analysis, molar conductance, thermogravimatric analysis and chemistry analysis. The standard molar enthalpies of solution of LaCl3•7H2O (s), [2C7H6O3 (s)], C4H7NO2S (s) and [La(Hsal)2•(tch)]•2H2O (s) in a mixed solvent of absolute ethyl alcohol, dimethyl sulfoxide (DMSO) and 3 mol•L-1 HCl were determined by calorimetry to be [LaCl3•7H2O (s), 298.15 K]=(-102.36±0.66) kJ•mol-1, [2C7H6O3 (s), 298.15 K]=(26.65±0.22) kJ•mol-1, [C4H7NO2S (s), 298.15 K]=(-21.79±0.35) kJ•mol-1 and {[La(Hsal)2•(tch)]•2H2O (s), 298.15 K}=(-41.10±0.32) kJ•mol-1. The enthalpy change of the reaction LaCl3•7H2O (s)+2C7H6O3 (s)+C4H7NO2S (s)=[La(Hsal)2•(tch)]•2H2O (s)+3HCl (g)+5H2O (l) (Eq. 1) was determined to be =(41.02±0.85) kJ•mol-1. From date in the literature, through Hess’ law, the standard molar enthalpy of formation of [La(Hsal)2•(tch)]•2H2O (s) was estimated to be {[La(Hsal)2•(tch)]•2H2O (s), 298.15 K}=(-3017.0±3.7) kJ•mol-1.  相似文献   

15.
Complexes [{Ru(CO)Cl(PiPr3)2}2(μ‐2,5‐(CH?CH)2cC4H2E] (E=NR; R=C6H4‐4‐NMe2 ( 10 a ), C6H4‐4‐OMe ( 10 b ), C6H4‐4‐Me ( 10 c ), C6H5 ( 10 d ), C6H4‐4‐CO2Et ( 10 e ), C6H4‐4‐NO2 ( 10 f ), C6H3‐3,5‐(CF3)2 ( 10 g ), CH3 ( 11 ); E=O ( 12 ), S ( 13 )) are discussed. The solid state structures of four alkynes and two complexes are reported. (Spectro)electrochemical studies show a moderate influence of the nature of the heteroatom and the electron‐donating or ‐withdrawing substituents R in 10 a – g on the electrochemical and spectroscopic properties. The CVs display two consecutive one‐electron redox events with ΔE°′=350–495 mV. A linear relationship between ΔE°′ and the σp Hammett constant for 10 a–f was found. IR, UV/Vis/NIR and EPR studies for 10 +– 13 + confirm full charge delocalization over the {Ru}CH?CH‐heterocycle‐CH?CH{Ru} backbone, classifying them as Class III systems according to the Robin and Day classification. DFT‐optimized structures of the neutral complexes agree well with the experimental ones and provide insight into the structural consequences of stepwise oxidations.  相似文献   

16.
An examination of the results of measurements of the forward and reverse rate constants for the reaction shows that agreement between the kinetics and the thermochemistry is achieved only through use of a value of ΔHf(C2H5) = 28 kcal mol?1. This system therefore provides further support for the recent measurement of this quantity.  相似文献   

17.
The kinetics of formation and dissociation of [V(H2O)5NCS]2+ have been studied, as a function of excess metal-ion concentration, temperature, and pressure, by the stopped-flow technique. The thermodynamic stability of the complex was also determined spectrophotometrically. The kinetic and equilibrium data were submitted to a combined analysis. The rate constants and activation parameters for the formation (f) and dissociation (r) of the complex are: k/M ?1 · S?1 = 126.4, k/s?1 = 0.82; ΔH /kJ · mol?1 = 49.1, ΔH/kJ · mol?1 = 60.6; ΔS/ J·K?1·mol?1= ?39.8, ΔSJ·K?1·mol?1 = ?43.4; ΔV/cm3·mol?1 = ?9.4, and ΔV/cm3 · mol?1 =?17.9. The equilibrium constant for the formation of the monoisothiocynato complex is K298/M ?1 = 152.9, and the enthalpy and entropy of reaction are ΔH0/kJ · mol?1 = ? 11.4 and ΔS0/J. K?1mol?1 = +3.6. The reaction volume is ΔV0/cm3· mol?1 = +8.5. The activation parameters for the complex-formation step are similar to those for the water exchange on [V(H2O)6]3+ obtained previously by NMR techniques. The activation volumes for the two processes are consistent with an associative interchange, Ia, mechanism.  相似文献   

18.
Using published data on the kinetics of pyrolysis of C2Cl6 and estimated rate parameters for all the involved radical reactions, a mechanism is proposed which accounts quantitatively for all the observations: The steady-state rate law valid for after about 0.1% reaction is and the reaction is verified to proceed through the two parallel stages suggested earlier whose net reaction is A reported induction period obtained from pressure measurements used to follow the rate is shown to be compatible with the endothermicity of reaction A, giving rise to a self-cooling of the gaseous mixture and thus an overall pressure decrease. From the analysis, the bond dissociation energy DH0(C2Cl5? Cl) is found to be 70.3 ± 1 kcal/mol and ΔHf3000(·C2Cl5) = 7.7 ± 1 kcal/mol. The resulting π? bond energy in C2Cl4 is 52.5 ± 1 kcal/mol.  相似文献   

19.
Thermodynamic properties (ΔH°f(298), S°(298) and Cp(T) from 300 to 1500 K) for reactants, adducts, transition states, and products in reactions of CH3 and C2H5 with Cl2 are calculated using CBSQ//MP2/6‐311G(d,p). Molecular structures and vibration frequencies are determined at the MP2/6‐311G(d,p), with single‐point calculations for energy at QCISD(T)/6‐311 + G(d,p), MP4(SDQ)/CbsB4, and MP2/CBSB3 levels of calculation with scaled vibration frequencies. Contributions of rotational frequencies for S°(298) and Cp(T)'s are calculated based on rotational barrier heights and moments of inertia using the method of Pitzer and Gwinn [1]. Thermodynamic parameters, ΔH°f(298), S°(298), and CP(T), are evaluated for C1 and C2 chlorocarbon molecules and radicals. These thermodynamic properties are used in evaluation and comparison of Cl2 + R· → Cl· + RCl (defined forward direction) reaction rate constants from the kinetics literature for comparison with the calculations. Data from some 20 reactions in the literature show linearity on a plot of Eafwd vs. ΔHrxn,fwd, yielding a slope of (0.38 ± 0.04) and intercept of (10.12 ± 0.81) kcal/mole. A correlation of average Arrhenius preexponential factor for Cl· + RCl → Cl2 + R· (reverse rxn) of (4.44 ± 1.58) × 1013 cm3/mol‐sec on a per‐chlorine basis is obtained with EaRev = (0.64 ± 0.04) × ΔHrxn,Rev + (9.72 ± 0.83) kcal/mole, where EaRev is 0.0 if ΔHrxn,Rev is more than 15.2 kcal/mole exothermic. Kinetic evaluations of literature data are also performed for classes of reactions. Eafwd = (0.39 ± 0.11) × ΔHrxn,fwd + (10.49 ± 2.21) kcal/mole and average Afwd = (5.89 ± 2.48) × 1012 cm3/mole‐sec for hydrocarbons: Eafwd = (0.40 ± 0.07) × ΔHrxn,fwd + (10.32 ± 1.31) kcal/mole and average Afwd = (6.89 ± 2.15) × 1011 cm3/mole‐sec for C1 chlorocarbons: Eafwd = (0.33 ± 0.08) × ΔHrxn,fwd + (9.46 ± 1.35) kcal/mole and average Afwd = (4.64 ± 2.10) × 1011 cm3/mole‐sec for C2 chlorocarbons. Calculation results on the methyl and ethyl reactions with Cl2 show agreement with the experimental data after an adjustment of +2.3 kcal/mole is made in the calculated negative Ea's. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 548–565, 2000  相似文献   

20.
The synthesis of a novel series of the intermediates N2(N3)‐[1‐alkyl(aryl/heteroaryl)‐3‐oxo‐4,4,4‐trifluoroalk‐1‐en‐1‐yl]‐2‐aminopyridines [F3CC(O)CH?CR1(2? NH?C5H3N)] and 2,3‐diaminopyridines [F3CC(O)CH?CR1(2‐NH2‐3‐NH? C5H3N)], where R1 = H, Me, C6H5, 4‐FC6H4, 4‐CIC6H4, 4‐BrC6H4, 4‐CH3C6H4, 4‐OCH3C6H4, 4,4′‐biphenyl, 1‐naphthyl, 2‐thienyl, 2‐furyl, is reported. The corresponding series of 2‐aryl(heteroaryl)‐4‐trifluoromethyl‐3H‐pyrido[2,3‐b][1,4]diazepin‐4‐ols obtained from intramolecular cyclization reaction of the respective trifluoroacetyl enamines or from the direct cyclocondensation reaction of 4‐methoxy‐1,1,1‐trifluoroalk‐3‐en‐2‐ones with 2,3‐diaminopyridine, under mild conditions, is also reported.  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