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 78 毫秒
1.
The far-UV (193 nm) laser flash photolysis of nitrogen-saturated isooctane solutions of 1,1-dimethylsiletane allows the direct detection of 1,1-dimethylsilene as a transient species, which (at low laser intensities) decays with pseudo-first-order kinetics (τ 10 μs) and exhibits a UV absorption spectrum with λmax 255 nm. Characteristic rapid quenching is observed for the silene with methanol (kMcOH = (4.9 ± 0.2) × 109 M−1 s−1), tert-butanol (kBuOH = (1.8 ± 0.1) × 109 M−1 s−1) and oxygen (kO2 = (2.0 ± 0.5) × 108 M−1 s−1). The Arrhenius activation parameters for the reaction with methanol have been determined to be Ea = −2.6 ± 0.6 kcal mol−1 and log A = 7.7 ± 0.3.  相似文献   

2.
The Arrhenius equation corresponding to the process P---Ag+P*---Ag*→---P---Ag*+P*---Ag has been determined for [(η6-p-cymene)Ru(μ-pz)3Ag(PPh3)] (1) by complete line-shape analysis of the 31P NMR spectra between −40°C and +30°C. It has the form K = 1011.8± e(−46±5 kJ mol−1/RT). The preexponential term, log A = 11.8 corresponds to a small activation entropy, whereas the activation energy, 46 kJ mol−1 is comparable to those determined for other phosphorus—metal compounds.  相似文献   

3.
Thermal decomposition of poly(1,4-dioxan-2-one)   总被引:2,自引:0,他引:2  
To evaluate the feasibility of poly(1,4-dioxan-2-one) (PPDO) as a feed stock recycling material, the pyrolysis kinetics of PPDO were investigated. The pyrolysis of PPDO exclusively resulted in the distillation of 1,4-dioxan-2-one (PDO). From thermogravimetric measurements conducted at different heating rates, the kinetic parameters of the pyrolysis: activation energy, Ea=127 kJ mol−1; order of reaction, n=0; and pre-exponential factor, A=2.3×109 s−1, were estimated by plural analytical methods. The estimates show that the decomposition of PPDO proceeds by unzipping depolymerization as main reaction and random degradation process with lower Ea and A values. Equivalent isothermal degradation curves calculated from the thermogravimetric curves were supported by experimental isothermal degradation data. The calculation that PPDO is converted smoothly into PDO at 270°C agrees with the reported ceiling temperature of PPDO.  相似文献   

4.
The hydration energy of metallic cations determined with density functional calculations using a double-numerical plus p-polarization basis set, related to the acidity constants of hexaaqua metal complexes, was investigated in the present study. From the results calculated by Vosko-Wilk-Nusair (VWN), Becke-Perdew (BP) and Becke-Lee-Yang-Parr (BLYP) density functionals, a global linear correlation with the observed acidity constants in both main group [Mg(II), Ca(II) and Al(III)] and (post-)transition group [Mn(II), Zn(II), Cd(II), Sc(III), Cr(III), Fe(III), Ga(III) and In(III)] hexaaqua metal complexes has been established:

VWN density functional: pKa = 16.5760 + 0.0173Ehydr kcal mol−1

BP density functional: pKa = 15.7329 + 0.0182Ehydr kcal mol−1

BLYP density functional: pKa = 15.9448 + 0.0185Ehydr kcal mol−1  相似文献   


5.
The oxidation reaction of 2-aminophenol (OAP) to 2-aminophenoxazin-3-one (APX) initiated by 2,2,6,6-tetramethyl-1-piperidinyloxyl (TEMPO) has been investigated in methanol at ambient temperature. The oxidation of OAP was followed by electronic spectroscopy and the rate constants were determined according to the rate law −d[OAP]/dt=kobs[OAP][TEMPO]. The rate constant, activation enthalpy and entropy at 298 K are as follows: kobs (dm3 mol−1 s−1)=(1.49±0.02)×10−4, Ea=18±5 kJ mol−1, ΔH=15±4 kJ mol−1, ΔS=−82±17 J mol−1 K−1. The results of oxidation of OAP show that the formation of 2-aminophenoxyl radical is the key step in the activation process of the substrate.  相似文献   

6.
Treatment of CpZrCl3 with 3-methylbutenyl-Grignard reagent yields thermally labile tris(1,1-dimethylallyl) ZrCp (6), which is slowly decomposed (5d) at −15°C to give (η-cyclopentadienyl)(η3-1,1-dimethylallyl)(η4-isoprene)zirconium (7), which is thermally unstable; with a half-live of 43 h at 20°C it rearranges to the η3-1,2-dimethylallyl isomer and an (isoprene) zirconium hydride is proposed as the intermediate for this hydrogen-migration reaction.  相似文献   

7.
Monohydrated sodium carbonate crystals have been grown by slow evaporation of its aqueous solution maintained at 40 ± 1°C. The thermal dehydration of this crystal has been studied by dynamic and isothermal TG measurements. It is observed from dynamic TG that the single molecule of water of crystallization is lost in two steps of 0.3 mole and 0.7 mole at temperatures 426 ± 5 and 454 ± 5 K, respectively. From isothermal and dynamic TG measurements, the kinetic parameters E and Z are calculated using different known forms of the function F(). It is observed that consistency of E and Z values in isothermal and dynamic TG measurements for the two dehydration steps gives the correct function F() = −[log(1-)]0.5. The activation energies for this function for the two dehydration steps are ≈6 and ≈9 kcal mole−1, respectively.  相似文献   

8.
The kinetics of the thermal decomposition of CoOOH powder has been studied isothermally in a temperature range of 260—310°C in air. The reaction was found to proceed by the advance of a two-dimensional reaction interface. The kinetics results indicate that there are two phases in the decomposition in this temperature range: up to 280°C with an activation energy E1 = 34.75 kcal mol−1 and above 280°C with E2 = 18.91 kcal mol−1. A reaction mechanism is proposed to account for these observations.  相似文献   

9.
The thermal decomposition and thermal stability of 1,3,5-trinitro-2-oxo-1,3,5-triazacyclohexane (keto-RDX or K-6) was studied. The keto-RDX synthesis is described, mass spectra (electron impact (70 eV) and chemical ionization) similar to RDX spectra registered under identical conditions are presented, and mass spectroscopy fragmentation paths are proposed. The LI-MS (laser induced/mass spectroscopic) results imply that the first step in the decomposition of keto-RDX is the elimination of NO2 or HONO and subsequent breakdown of the triazacyclohexane ring. The thermal stability, activation energy (Ea = 140 kJ mol−1), and frequency factor (K0 = 9 × 109 s−1) in the temperature interval 90-120°C were measured using chemiluminescence (NO detection only). The activation energy was also determined from DSC data using the ASTM method E 698-79, and was found to be 280 kJ mol−1 with a frequency factor of 7.0 × 1029 s−1 in the temperature interval 175-200°C. Microcalorimetry, drop-weight test, friction test, and ignition temperature (Wood's metal bath) measurements were also conducted. Quantum mechanical calculations (semi-empirical method with PM3 set at the unrestricted Hartree-Fock level) were conducted to correlate the sensitivity and thermal decomposition with those of RDX. No significant differences in bond-breaking energies for RDX and keto-RDX were found. Conclusions drawn from the experiments are that the decomposition of keto-RDX is auto-catalytic, and that the sensitivity of keto-RDX is not connected with the initial bond-breaking step. More than one method for measuring the risk involved in handling an explosive is necessary since the sensitivity depends on different stages in the decomposition.  相似文献   

10.
Atomization of germanium from zirconium-coated (ZrGT), palladium-coated (PdGT) and palladium-zirconium-coated graphite tubes (PZGT) by aqueous deposition of the analyte solution and/or the trapping of the gaseous hydride has been investigated. From the activation energies calculated based on Smets' method, it was found that both the mode of sample introduction and the nature of the atomizer surface have an effect on the atom formation of germanium. Activation energies Ea of 383.9 ± 17.1, 445.8 ± 19.8 and 557.9 ± 12.4 kJ mol−1 were observed for germanium atomization from ZrGT, PZGT and PdGT, respectively. A much larger Ea value of 950.2 ± 13.1 kJ mol−1 was obtained for Ge from PdGT by the trapping of GeH4. XPS results do not give sufficient evidence for Pd-Ge compound formation.  相似文献   

11.
The molecular structure of 1,1,2,2-tetrabromodisilane has been investigated using gas-phase electron diffraction data obtained at 110°C. At this temperature the molecules exist as a mixture of about equal parts (X = 0.5 ±0.2) of the two conformers with the H---Si---Si---H torsion angle equal to 180° (anti) or 60° (gauche). Assuming that the two conformers differ in their geometries only in the torsion angle φ, some of the important distance (ra) and angle () parameters are: r(Si---Si) = 2.349(19) Å, r(Si---Br) = 2.205(5) Å, r(Si---H) = 1.485 Å (assumed), Br---Si---Br = 110.1(1.6)°, Si---Si---Br = 107.1(1.2)° Si---Si---H = 108.6° (assumed). The error limits are 2σ. The observed conformational composition (Xanti = 0.5(0.2)) corresponds to an energy difference between the conformers of ΔE = E(gauche) — E(anti) = 0.5 ± 0.6 kcal mol−1, assuming ΔS = Rln2.  相似文献   

12.
The enthalpy of formation (ΔHf0), enthalpy of evaporation (ΔHv0) and enthalpy of atomization (ΔHa) of permethylcyclosilazanes (Me2SiNH)n (n = 3, 4) and 1,1,3,3-tetramethyldisilazane (Me2SiH)2NH have been determined. The enthalpies of formation of these compounds were compared with those calculated by the Benson-Buss-Franklin and Tatevskii additive schemes. In higher permethylcyclosilazanes the energy of the endocyclic Si---N bond is 306 ± 2 kJ mol−1 (73 kcal mol−1), that is 12 ± 2 kJ mol−1 (3 kcal mol−1) lower than the energy of the acyclic Si---N bond. The strain energy of the cyclotrisilazane ring is estimated to be 10.5 kJ mol−1 (2.5 kcal mol−1), whereas the energy of the ring Si---N bond is 295 kJ mol−1 (70.5 kcal mol−1).

The thermochemical data for permethylcyclosilazanes were compared with the corresponding values for permethylcyclosiloxanes calculated from the results of previously reported studies.  相似文献   


13.
The compound [RU332- -ampy)(μ3η12-PhC=CHPh)(CO)6(PPh3)2] (1) (ampy = 2-amino-6-methylpyridinate) has been prepared by reaction of [RU3(η-H)(μ32- ampy) (μ,η12-PhC=CHPh)(CO)7(PPh3)] with triphenylphosphine at room temperature. However, the reaction of [RU3(μ-H)(μ3, η2 -ampy)(CO)7(PPh3)2] with diphenylacetylene requires a higher temperature (110°C) and does not give complex 1 but the phenyl derivative [RU332-ampy)(μ,η 12 -PhC=CHPh)(μ,-PPh2)(Ph)(CO)5(PPh3)] (2). The thermolysis of complex 1 (110°C) also gives complex 2 quantitatively. Both 1 and 2 have been characterized by0 X-ray diffraction methods. Complex 1 is a catalyst precursor for the homogeneous hydrogenation of diphenylacetylene to a mixture of cis- and trans -stilbene under mild conditions (80°C, 1 atm. of H2), although progressive deactivation of the catalytic species is observed. The dihydride [RU3(μ-H)232-ampy)(μ,η12- PhC=CHPh)(CO)5(PPh3)2] (3), which has been characterized spectroscopically, is an intermediate in the catalytic hydrogenation reaction.  相似文献   

14.
Pentacarbonyl(diethylaminocarbyne)chromium tetrafluoroborate, [(CO)5− CrCNEt2]BF4 (I), reacts with PPh3 with substitution of CO and formation of trans-tetracarbonyl(diethylaminocarbyne)triphenylphosphanechromium tetra-fluoroborate, trans-[PPh3(CO)4CrCNEt2]BF4 (III). Substitution of CO by PPh3 in neutral trans-tetracarbonyl(halo)(diethylaminocarbyne)chromium complexes, trans-X(CO)4CrCNEt2 (IVa: X = Br, IVb: X = I), leads in a reversible reaction to the corresponding tricarbonyl complexes, mer-X(PPh3)(CO)3− CrNEt2 (V), PPh3 occupying the cis-position to the carbyne ligand. With PPh3 in large excess both reactions follow a first-order rate law. This as well as the activation parameters (ΔH≠ = 104–113 kJ mol−1, ΔS≠ = 64–71 J mol−1 K−1) indicate a dissociative mechanism.  相似文献   

15.
The butadienyl complexes formed by the reaction of trans-(R1)CH=CHCCR2 (R1, R2 = SiMe3, tBu, Me, Et) with RuCl(CO)H(PPh3)3 exhibit unique structures: instead of taking the 18-electron configuration of the metal by conventional η3-coordination of the butadienyl ligand, they shift significantly to the 16-electron η1-coordination state.  相似文献   

16.
UV irradiation of tricarbonyl-η5-2,4-dimethyl-2,4-pentadien-1-yl-manganese (2) in THF at 208 K yields solvent-stabilized dicarbonyl-η5-2,4-dimethyl-2,4-pentadien-1-yl-tetrahydrofurane-manganese (3), which reacts in situ with two equivalents of 1-dimethylamino-2-propyne (4) to dicarbonyl-1–5-η-2,4-dimethyl-(6-dimethylaminomethyl-N)-10-dimethylamino-deca-2,4,6,8- tetraen-1-yl-manganese (5). The crystal and molecular structure was determined by an X-ray diffraction analysis. Complex 5 crystallizes in the monoclinic space group P21/c, A = 1109.9(2) pm, B = 836.0(2) pm, C = 2156.9(4) pm, β = 93.23(3)°, V = 1.9982(7) nm3, Z = 4. Complex 5 was also studied in solution by IR and NMR spectroscopy. A possible formation mechanism of 5 will be discussed.

Zusammenfassung

UV-Bestrahlung von Tricarbonyl-η5-2,4-dimethyl-2,4-pentadien-1-yl-mangan (2) in THF bei 208 K liefert solvenstabilisiertes Dicarbonyl-η5-2,4-dimethyl-2, 4-pentadien-1-yl-tetrahydrofuran-mangan (3), welches in situ mit zwei Äquivalenten 1-Dimethylamino-2-propin (4) zu Dicarbonyl-1–5-η-2,4-dimethyl-(6-dimethylaminomethyl-N)-10-dimethylamino-deca-2,4,6,8-tetraen-1-yl-mangan (5) reagiert. Seine Kristall- und Molekülstruktur wurde durch eine Röntgenbeugungsanalye bestimmt. Komplex 5 kristallisiert in der monoclinen Raumgruppe P21/c, A = 1109.9(2) pm, B = 836.0(2) pm, C = 2156.9(4) pm, β = 93.23(3)°, V = 1.9982(7)_ nm3, Z = 4. Komplex 5 wurde auch in Lösung IR- und NMR-spektroskopisch untersucht. Ein möglicher Bildungsmechanismus von 5 wird diskutiert.  相似文献   


17.
The reactions of hydroxyl radical, hydrogen atom and hydrated electron intermediates of water radiolysis with N-isopropylacrylamide (NIPAAm) were studied by pulse radiolysis in dilute aqueous solutions. OH, H and eaq react with NIPAAm with rate coefficient of (6.9±1.2)×109, (6.6±1)×109, and (1.0±0.2)×1010 mol−1 dm3 s−1. In OH and H radical addition to the double bond mainly -carboxyalkyl type radicals form, (OHCH2CHC(N-i-C3H7)O and CH3CHC(N-i-C3H7)O). In reaction of eaq oxygen atom centered radical anion is produced (CH2CHC(N-i-C3H7)O), the anion undergoes reversible protonation with pKa=8.7. There is also an irreversible protonation on the β-carbon atom that produces the same radical as forms in H atom reaction (CH3CHC(N-i-C3H7)O). The -carboxyalkyl type radicals at low NIPAAm concentration (0.1–1 mmol dm−3) mainly disappear in self-termination reactions, 2kt,m=8.4×108 mol−1 dm3 s−1. At higher concentrations the decay curves reflect the competition of the self-termination and radical addition to monomer (propagation). The termination rate coefficient of oligomer radicals containing a few monomer units is 2kt≈2×108 mol−1 dm3 s1.  相似文献   

18.
Gaussian-2 ab initio calculations were performed to examine the six modes of unimolecular dissociation of cis-CH3CHSH+ (1+), trans-CH3CHSH+ (2+), and CH3SCH2+ (3+): 1+→CH3++trans-HCSH (1); 1+→CH3+trans-HCSH+ (2); 1+→CH4+HCS+ (3); 1+→H2+c-CH2CHS+ (4); 2+→H2+CH3CS+ (5); and 3+→H2+c-CH2CHS+ (6). Reactions (1) and (2) have endothermicities of 584 and 496 kJ mol−1, respectively. Loss of CH4 from 1+ (reaction (3)) proceeds through proton transfer from the S atom to the methyl group, followed by cleavage of the C–C bond. The reaction pathway has an energy barrier of 292 kJ mol−1 and a transition state with a wide spectrum of nonclassical structures. Reaction (4) has a critical energy of 296 kJ mol−1 and it also proceeds through the same proton transfer step as reaction (3), followed by elimination of H2. Formation of CH3CS+ from 2+ (reaction (5)) by loss of H2 proceeds through protonation of the methine (CH) group, followed by dissociation of the H2 moiety. Its energy barrier is 276 kJ mol−1. On both the MP2/6-31G* and QCISD/6-31G* potential-energy surfaces, the H2 1,1-elimination from 3+ (reaction (6)) proceeds via a nonclassical intermediate resembling c-CH3SCH2+ and has a critical energy of 269 kJ mol−1.  相似文献   

19.
The molecular and crystal structure of the nido-6-tungstadecaborane [6,6,6,6-(CO)2(PPh3)2-nido-6-WB9H13] (1) has been determined showing that the tungsten atom is incorporated into the 6-position of a nido 10-vertex (WB9) cage. The tungsten atom has a seven-coordinate capped trigonal prismatic environment and is bonded to two hydrogen and three boron atoms of the {B9H13} cage, in addition to two CO groups and two PPh3 ligands. Variable-temperature (−90°C to +50°C) 31P{1H} NMR spectroscopy of 1 reveals that the exo-polyhedral ligands about the tungsten atom are fluxional with respect to PPh3 site exchange with an activation energy (ΔG‡), at the coalescence temperature (−73°C), of <38 kJ mol−1.  相似文献   

20.
Polarized absorption spectra of Ba(MnO4)2·3H2O/Ba(ClO4)2·3H2O mixed single crystals are reported at 4.2°K. Previous 1T21A1 assignments for the 5200 Å and 3000 Å absorption bands of MnO4 are substantiated; further support is provided for the 1T11A1 assignment of the 3600 Å absorption band of MnO4. The site-splitting of the 5200 Å 1T2 state is E(1E)−E(1A) ≈ −150 cm−1; that of the 3000 Å 1T2 state is E(1E)−E(1A) ≈ 300 cm−1. A significant e vibronic intensity component is observed in the 5200 Å 1T2 state.  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