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 31 毫秒
1.
Changes in the magnitude of 〈ΔEdown〉, the average downward collisional energy transferred between a highly vibrationally excited reactant molecule and an inert bath gas, upon perdeuteration of the substrate are reported for tert-butyl bromide dilute in Ar, Kr, N2, and CO2. The technique of pressure-dependent very low-pressure pyrolysis (VLPP) was used to obtain the absolute values of 〈ΔEdown〉, which are for C4H9Br, 230 (Ar), 285 (Kr), 270 (N2), and 365 (CO2) while for C4D9Br, 200 (Ar), 250 (Kr), 220 (N2), and 335 (CO2), all in cm?1 at ca. 720 K. The estimated uncertainties in these values are ca. ± 10%. These observed 〈ΔEdown〉, values and trends found with results from this series of isotope studies, are compared with current theoretical models. Extrapolated high-pressure temperature-dependent rate coefficients (s?1) for the thermal decomposition of reactant are 1013.8±0.3 exp(?175 ± 8 kJ mol?1/RT) for C4H9Br and 1014.3±0.3 exp(?183 ± 8 kJ mol?1/RT) for C4D9Br. These results are in accord with other studies and the expected isotope effect.  相似文献   

2.
The average downward energy transfer (〈Δ Edown〉) is obtained for highly vibrationally excited acetyl chloride with Ne and C2H4 bath gases at ca. 870 K. Data are obtained by the technique of very low-pressure pyrolysis (VLPP). Fitting these data by solution of the appropriate reaction-diffusion integrodifferential master equation yields the gas/gas collisional energy transfer parameters: 〈Δ Edown〉 values are 220 ± 10 cm?1 (Ne bath gas) and 330 ± 20 cm?1 (C2H4). These energy transfer quantities are much less than those predicted by statistical theories, or those observed for similar sized molecules such as CH3CH2Cl. These results are explained by the qualitative predictions of the biased random walk model wherein the fundamental mechanism of energy transfer is the multiple interactions between the bath gas and the individual atoms of the reactant molecule, during the course of the collision event. The charge distribution of acetyl chloride decreases the number of such interactions, thereby reducing the amount of energy transferred per collision.  相似文献   

3.
The average downward collisional energy transfer (<ΔEdown>) is obtained for highly vibrationally excited tert-butyl chloride, both undeuterated and per-deuterated, with Kr, N2, CO2, and C2H4 bath gases, at ca. 760 K. Data are obtained using the technique of pressure-dependent very low-pressure pyrolysis. Reactant internal energies to which the data are sensitive are in the range 200–250 kJ mol?1. For C4H9Cl, the <ΔEdown> values (cm?1) are 255 (Kr), 265 (N2), 440 (CO2), and 585 (C2H4), and for C4D9Cl, 245 (N2), 370 (CO2), and 540 (C2H4). The uncertainties in these values are ca. 20% (40% for Kr); the uncertainties in the deuteration ratios are 10–15%. The value for Kr is in agreement with theoretical predictions of a biased random walk model for internal energy change in monatomic/substrate collisions. The effect of deuteration of <ΔEdown> is also in accord with that predicted by a modification of the theory. Extrapolated highpressure rate coefficients for the thermal decomposition of reactant are 1013.6 exp(-187 kJ mol?1/RT) s?1 (C4H9Cl) and 1014.2 exp(?196 kJ mol?1/RT) s?1 (C4D9Cl), in accord with other studies and the expected isotope effect.  相似文献   

4.
Br-atom atomic resonance absorption spectrometry (ARAS) has been developed and applied to measure thermal decomposition rate constants for CF3Br (+ Kr)→CF3+Br (+ Kr) over the temperature range, 1222–1624 K. The Br-atom curve-of-growth (145<λ<163 nm) was determined using this reaction. For [Br]≤1×1012 molecules cm−3, absorbance, (ABS)=1.410×10−13 [Br], yielding σ=1.419×10−14 cm2. The curve-of-growth was then used to convert (ABS) to Br-atom profiles which were then analyzed to give measured rate constants. These can be expressed in second-order by k1=8.147×10−9 exp(−24488 K/T) cm3 molecule−1 s−1 (±33%, 1222≤T≤1624 K). A unimolecular theoretical approach was used to rationalize the data. Theory indicates that the dissociation rates are closer to second- than to first-order, i.e., the magnitudes are 30–53% of the low-pressure-limit rate constants over 1222–1624 K and 123–757 torr. With the known, E0=ΔH00=70.1 kcal mole−1, the optimized theoretical fit to the ARAS data requires 〈ΔEdown=550 cm−1. These conclusions are consistent with recently published data and theory from Kiefer and Sathyanarayana. © 1998 John Wiley & Sons, Inc. Int J Chem Kinet 30: 859–867, 1998  相似文献   

5.
6.
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   

7.
Vinylacetylene was pyrolyzed at 300–450°C in a packed and an unpacked static reactor with a pinhole bleed to a quadrupole mass spectrometer. The reactant and C8H8 products were monitored continuously during a reaction by mass spectrometry. In some runs, the products were also analyzed by gas chromatography after the run. In these runs CH4, C2H6, C3H6, and C2H4 were also detected. The reaction for vinylacetylene removal and C8H8 formation is homogeneous, second order in reactant, and independent of the presence of a large excess of N2 or He. However, C8H8 formation is about half-suppressed by the addition of the free-radical scavengers NO or O2. The rate coefficient for total vinylacetylene removal is 1.7 × 106 exp(?79 ± 13 kJ/mol RT) L/mol · s. The major reaction for C4H4 removal is polymerization. In addition four C8H8 isomers, carbon, and small hydrocarbons are formed. The three major C8H8 isomers are styrene, cyclooctatetraene (COT), and 1,5? dihydropentalene (DHP). The C8H8 compounds are formed by both molecular and free-radical processes in a second-order process with an overall k ? 3 × 108 exp(?122 kJ/mol RT) L/mol · s (average of packed and unpacked cell results). The molecular process occurs with an overall k = 8.5 × 107 exp (?118 kJ/mol RT) L/mol · s. The COT, DHP, and an unidentified isomer (d), are formed exclusively in molecular processes with respective rate coefficients of 4.4 × 104 exp(?77 kJ/mol RT), 1.7 × 105 exp(?89 kJ/mol RT), and 3.1 × 109 exp(? 148 kJ/mol RT) L/mol · s. The styrene is formed both by a direct free-radical process and by isomerization of COT.  相似文献   

8.
The rate constants 〈σ · υ〉 for collisional de-excitation of the metastable 5D states of Ba+ ions have been determined in an ion trap experiment. TheD-states are selectively populated by pulsed laser excitation of the 6P 1/2 or 6P 3/2 state and the decay at different background pressures is monitored by the change in fluorescence intensity of the excited ions. From the pressure dependence of the decay constants we calculate the de-excitation rate constants for different collision partners, averaged over the velocity distribution of the trapped ion cloud. For He, Ne, H2 and N2 we obtain in the c.m. energy range of 0.1–0.5 eV: 〈σ·υ〉 (He)=3.0±0.2·10?13cm3/s, 〈σ·υ〉 (Ne)=5.1±0.4·10?13cm3/s, 〈σ·υ〉 (H2)=3.7±0.3·10?11cm3/s, 〈σ·υ〉 (N2)=4.4±0.3·10?11cm3/s. The results can be understood qualitatively by a consideration of the ion-atom and ion-molecules interaction potential.  相似文献   

9.
A detailed master equation simulation has been carried out for the thermal unimolecular decomposition of C6H10 in a shock tube. At the highest temperatures studied experimentally [J. H. Kiefer and J. N. Shah, J. Phys. Chem., 91, 3024 (1987)], the average thermal vibrational energy is greater than the reaction threshold and therefore 〈ΔE〉 (up and down steps) is positive for molecules at that energy, rather than negative; the converse is true at lower temperatures. The calculated incubation time, in which the decomposition rate constant rises to 1/e of its steady state value, is found to be only weakly dependent on temperature (at constant pressure) between 1500 K and 2000 K and to depend almost exclusively on 〈ΔEd (down steps, only), and not on collision probability model. Simulations of the experimental data show the magnitude of 〈ΔEd depends weakly on assumed collision probability model, but is nearly independent of temperature. The second moment 〈ΔE½ is found to be independent of both temperature and transition probability model. The experimental data are not very sensitive to the possible energy-dependence of 〈ΔEd for a wide range of assumptions. It is concluded that the observed experimental “delay times” probably can be identified with the incubation time; further experiments are desirable to test this possibility and obtain more direct measures of the incubation time.  相似文献   

10.
The reactions of ground-state oxygen atoms with carbonothioicdichloride, carbonothioicdifluoride, and tetrafluoro-1,3-dithietane have been studied in a crossed molecular jet reactor in order to determine the initial reaction products and in a fast-flow reactor in order to determine their overall rate constants at temperatures between 250 and 500 K. These rate constants are??(O + C2CS) =(3.09 ± 0.54) × 10?11 exp(+115 ± 106 cal/mol/RT),??(O + F2CS) = (1.22 ± 0.19) × 10?11 exp(-747 ± 95 cal/mol/RT), and??(O + F4C2S2) = (2.36 ± 0.52) × 10?11 exp(-1700 ± 128 cal/mol/RT) cm3/molec˙sec. The detected reaction products and their rate constants indicate that the primary reaction mechanism is the electrophilic addition of the oxygen atom to the sulfur atom contained in the reactant molecule to form an energy-rich adduct which then decomposes by C-S bond cleavage.  相似文献   

11.
Laser flash photolysis coupled with resonance fluorescence detection of Br atoms was employed to investigate the temperature dependence of the reaction Br + neo‐C5H12 (1) between 688 and 775 K. The following Arrhenius preexponential factor and activation energy were determined (±1 σ): A1 = (6.89 ± 2.27) 1014 cm3 mol−1 s−1 and EA,1 = 57.61 ± 2.05 kJ mol1 The only other kinetic parameters reported for the reaction of Br atoms with neo‐C5H12 were obtained from competitive kinetic experiments relative to Br + C2H6. Comparison with our direct results is hampered by uncertainties in the kinetic data for the reference reaction that may need reinvestigation. The standard enthalpy of formation for the neo‐C5H11 radical was estimated to be 34.7 and 41.6 kJ mol−1, depending on the value of the activation energy assumed for the reverse reaction neo‐C5H11 + HBr (−1). © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 33: 49–55, 2001  相似文献   

12.
The thermal decomposition of 3,4-dihydro-2H-pyran (DHP, C5H8O) has been investigated by two methods: in shock waves with the laser-schlieren technique using mixtures of 5 and 10% DHP in krypton over 900–1500 K, 110–560 torr; in a flow tube having a reaction pressure 0.5 torr above atmospheric using the decomposition of allylethyl ether as an internal standard, and covering 663–773 K. The retro-Diels-Alder dissociation to the stable acrolein and ethylene is the dominant channel for all conditions. Precise rate constants (rms deviation of 10%) were obtained for this process over the indicated temperature ranges. Unimolecular falloff is evident in the shock-tube results, and RRKM calculations also predict a slight falloff at the lower temperatures. These RRKM calculations use a routine vibration model transition state and agree closely with the high-temperature data when 〈ΔEdown is a fixed 400 cm?1. Arrhenius expressions for k derived from the two measurements are in close accord and also consistent with most previous studies of this reaction. © 1995 John Wiley & Sons, Inc.  相似文献   

13.
The kinetic isotope effect for the abstraction of hydrogen/deuterium from dimethylnitramine and dimethylnitramine-d6 by chlorine atoms has been studied in the temperature range 273–353 K. The rate constant ratio kH0/kD is given by the Arrhenius expression, kH/kD=(0.92 ± 0.07)exp(286 ± 250/RT), where R is expressed in cal mol?1 K?1. The absolute rate constant for the deuterium abstraction reaction is extrapolated as kD=(1.50 ± 0.90) × 10?10 exp(?1,486 ± 370/RT) cm3 molecule?1 s?1. The temperature dependence of the kinetic isotope effect was calculated using the conventional transition-state theory, and the obtained values for kH/kD and ΔEH, D are in good agreement with the experimental value for a bent transition state geometry, with two new vibrational frequencies of 340 cm?1 (272 cm?1) corresponding to the in-plane and out-of-plane motions of hydrogen (deuterium) atoms in the Cl…H…C arrangement. © 1993 John Wiley & Sons, Inc.  相似文献   

14.
Reactions of ozone with simple olefins have been studied between 6 and 800 mtorr total pressure in a 220-m3 reactor. Rate constants for the removal of ozone by an excess of olefin in the presence of 150 mtorr oxygen were determined over the temperature range 280 to 360° K by continuous optical absorption measurements at 2537 Å. The technique was tested by measuring the rate constants k1 and k2 of the reactions (1) NO + O3 → NO2 + O2 and (2) NO2 + O3 rarr; NO3 + O2 which are known from the literature. The results for NO, NO2, C2H4, C3H6, 2-butene (mixture of the isomers), 1,3→butadiene, isobutene, and 1,1 -difluoro-ethylene are 1.7 × 10?1 4 (290°K), 3.24 × 10?17 (289°K), 1.2 × 10?1 4 exp (–4.95 ± 0.20/RT), 1.1 × 10?1 4 exp (–3.91 ± 0.20/RT), 0.94 × 10?1 4 exp ( –2.28 ± 0.15/RT), 5.45 ± 10?1 4 exp ( –5.33 ± 0.20/RT), 1.8 ×10?17 (283°K), and 8 × 10?20 cm3/molecule ·s(290°K). Productformation from the ozone–propylene reaction was studied by a mass spectrometric technique. The stoichiometry of the reaction is near unity in the presence of molecular oxygen.  相似文献   

15.
Gas phase slow combustion of (chloro)benzene in O2/N2 mixtures, and induced by addends such as tert butylhydroperoxide, cyclohexane, or methanol, leads to (chloro)-phenol as the only important aromatic product. Using C6H6/C6D6 mixtures, formation of phenol/perdeuterophenol was studied between 520–1080 K. The temperature dependence of this product ratio was found to obey the Arrhenius expression for the intermolecular isotope effect log kH/kD = ?0.14 ± 0.03 + (1240 ± 80)/2.303RT (R in cal/mol K). Essentially the same result was obtained for the intramolecular isotope effect, measuring the change in isomer distribution for the chlorophenols formed from p-deuterio-chlorobenzene versus those for chlorobenzene. These results are in accordance with H(D)-abstraction by ·OH, via a linear transition state, as the first and (relative) rate determining step. Whereas above 1000 K, at reduced pressure, the intramolecular isotope effect continues to prevail, C6H6/C6D6 do not show differences in rate of formation of C6H5OH/C6D5OH. Under these conditions, the only effective reaction of arene to phenol appears to be set in by addition of O(3P).  相似文献   

16.
We report quantitative calculations of stereomutation tunneling in the disulfane isotopomers H2S2, D2S2, and T2S2, which are chiral in their equilibrium geometry. The quasi‐adiabatic channel, quasi‐harmonic reaction path Hamiltonian approach used here treats stereomutation including all internal degrees of freedom. The torsional motion is handled as an anharmonic reaction coordinate in detail, whereas all the remaining degrees of freedom are taken into account approximately. We predict how stereomutation is catalyzed or inhibited by excitation of the various vibrational modes. The agreement of our theoretical results with spectroscopic data from the literature on H2S2 and D2S2 is excellent. We furthermore predict the influence of parity violation on stereomutation as characterized approximately by the ratio (ΔEpv/ΔE±) of the (local or vibrationally averaged) parity violating potential ΔEpv and the tunneling splittings ΔE± in the symmetrical case. This ratio is exceedingly small for the reference molecules H2O2 and D2O2, and still very small (2⋅10−6 cm−1) for H2S2, which, thus, all exhibit essentially parity conservation in the dynamics. However, for D2S2 it is ca. 0.002, and for T2S2 it is ca. 1, which seems to be the first case where such intermediate mixing through parity violation is quantitatively predicted for spectroscopically accessible molecules. The consequences for the spectroscopic detection of molecular parity violation are discussed briefly also in relation to other molecules.  相似文献   

17.
The kinetics and equilibrium of the gas-phase reaction of CH3CF2Br with I2 were studied spectrophotometrically from 581 to 662°K and determined to be consistent with the following mechanism: A least squares analysis of the kinetic data taken in the initial stages of reaction resulted in log k1 (M?1 · sec?1) = (11.0 ± 0.3) - (27.7 ± 0.8)/θ where θ = 2.303 RT kcal/mol. The error represents one standard deviation. The equilibrium data were subjected to a “third-law” analysis using entropies and heat capacities estimated from group additivity to derive ΔHr° (623°K) = 10.3 ± 0.2 kcal/mol and ΔHrr (298°K) = 10.2 ± 0.2 kcal/mol. The enthalpy change at 298°K was combined with relevant bond dissociation energies to yield DH°(CH3CF2 - Br) = 68.6 ± 1 kcal/mol which is in excellent agreement with the kinetic data assuming that E2 = 0 ± 1 kcal/mol, namely; DH°(CH3CF2 - Br) = 68.6 ± 1.3 kcal/mol. These data also lead to ΔHf°(CH3CF2Br, g, 298°K) = -119.7 ± 1.5 kcal/mol.  相似文献   

18.
合成了高氯酸镨和咪唑(C3H4N2), DL-α-丙氨酸(C3H7NO2)混配配合物晶体. 经傅立叶变换红外光谱、化学分析和元素分析确定其组成为[Pr(C3H7NO2)2(C3H4N2)(H2O)](ClO4)3. 使用具有恒温环境的溶解-反应量热计, 以2.0 mol•L-1 HCl为量热溶剂, 在T=(298.150±0.001) K时测定出化学反应PrCl3•6H2O(s)+2C3H7NO2(s)+C3H4N2(s)+3NaClO4(s)=[Pr(C3H7NO2)2(C3H4N2)(H2O)](ClO4)3(s)+3NaCl(s)+5H2O(1)的标准摩尔反应焓为ΔrHmө=(39.26±0.11) kJ•mol-1. 根据盖斯定律, 计算出配合物的标准摩尔生成焓为ΔfHmө{[Pr(C3H7NO2)2(C3H4N2)(H2O)](ClO4)3(s), 298.150 K}=(-2424.2±3.3) kJ•mol-1. 采用TG-DTG技术研究了配合物在流动高纯氮气(99.99%)气氛中的非等温热分解动力学, 运用微分法(Achar-Brindley-sharp和Kissinger法)和积分法(Satava-Sestak和Coats-Redfern法)对非等温动力学数据进行分析, 求得分解反应的表观活化能E=108.9 kJ•mol-1, 动力学方程式为dα/dt=2(5.90×108/3)(1-α)[-ln(1-α)]-1exp(-108.9×103/RT).  相似文献   

19.
The competitive reactions of Br atoms with CH4 and CD4 were studied over the temperature range of 562° to 637°K. Over this temperature interval, the kinetic isotope effect, kH/kD, varied from 3.05 to 2.47 for the reactions The rate constant ratio kH/kD, expressed in Arrhenius form, was found to equal (1.10 ± 0.05) exp (1030 ± 60/RT). A comparison is presented between the experimental result and the result obtained theoretically from absolute rate theory using the London-Eyring-Polanyi-Sato (LEPS) method of constructing the potential energy surface of the reaction. The agreement between theory and experiment is very poor, and this is believed to arise from the highly unsymmetrical nature of the potential energy surface involved in these reactions. A comparison is also presented between the kH/kD values obtained in the Br + CH4–CD4 experiments and the available data on the corresponding Cl + CH4–CD4 reactions.  相似文献   

20.
Dissociation, vibrational relaxation, and unimolecular incubation have all been observed in shock waves in isobutene with the laser‐schlieren technique. Experiments covered a wide range of high‐temperature conditions: 900–2300 K, and post‐incident shock pressures from 7 to 400 torr in 2, 5, and 10% mixtures with krypton. The surprising observation is that of vibrational relaxation, well resolved over the full temperature range. The resolved process is completely exponential, with relaxation times in the range 20–120 ns atm. Relaxation and dissociation are clearly separated for T > 1850 K, with estimated incubation times near 200 ns atm. Incubation is essential for modeling of the very low‐pressure decomposition. Modeling of gradients with a chain mechanism initiated by CH fission produces an excellent fit and accurate dissociation rates that show severe falloff. A restricted‐rotor, Gorin‐model RRKM analysis fits these rates quite well with the known bond‐energy as barrier and 〈ΔEdown = 680 cm?1. The extrapolated k is log k(s?1) = 19.187–0.865 log T ?87.337 (kcal/mol)/RT, in good agreement with previous work. © 2003 Wiley Periodicals, Inc. Int J Chem Kinet 35: 381–390, 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