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 328 毫秒
1.
The selective laser excitation and induced fluorescence observation technique has been used to study rotationally inelastic collisions of I2*(B 0u+, υ = 15,j) with I2, 3He, 4He, Ne, Ar, H2 and D2. For each collision partner, several initial rotational levels ranging from ji = 12 up to ji = 146 have been excited. For purely rotational transfer within the υ = 15 level, our data are perfectly consistent with energy sudden (eventually corrected) scaling laws. Thus, any thermally averaged rate constant, k(jijf), can be expressed as a function of the basis rate constants k(l → 0). Furthermore, these k(l → 0) are found to follow simple empirical fitting laws. Consequently any k(jijf) can be predicted given a set of two or three fitting parameters. Collisions with relatively heavy particles (I2, Ar and Ne) are well described by using the inverse power fitting law k(l → 0) = b[l(l+1)], where b = 1.7, 1.2 and 1.2×10?10 cm3 s?1 and γ = 1.08, 1.02 and 1.17 for I2*-Ne, I2*-Ar and I2*-I2 collisions respectively. For collisions with light particles (3He, 4He, H2 and D2), k(l → 0) shows a sharp decrease with l which can be accounted for by a hybrid power-exponential fitting law k(l → 0) = b[l(l+1)] exp[-l(l+1)/l* (l*+1)], where b = 0.84, 0.71, 2.77 and 2.78×10?10 cm3 s?1l+ = 20.6, 23.1, 18.8 and 31.4, and γ = 0.66, 0.66, 0.78 and 0.91 for I2*-3He, I2*-He, I2*-H2 and I2*-D2 collisions, respectively. We confirm that the rotational transfer dynamics in heavy molecules is mainly governed by angular momentum exchange.  相似文献   

2.
In this communication are presented exact quantum mechanical nonadiabatic electronic transition probabilities for the collinear reaction Ar+ + H2(vi = 0) → ArH+(vf) + H. The calculations were performed using a potential surface calculated by the DIM method. It is established that large probabilities (≈ 1.0) can be obtained only if there is enough translational energy to overcome a potential barrier formed due to the crossing between vi = 0 of the Ar+ + H2 system and vi = 2 of the Ar + H+2 system. The threshold for the reaction is found to be 0.06 eV.  相似文献   

3.
Intramolecular vibration—vibration energy transfer cross sections have been calculated for CO2(0001) + H2/D2 → CO2(1110) + H2/D2, → CO2(1000) + H2/D2, and → CO2(2000) + H2/D2 based on the mechanism that the energy mismatch is transferred to the translational motion. For CO2 + H2, the calculated cross section for CO2(0001) + H2 → CO2(1000) + H2 is in good agreement with experimental data. Cross sections for the processes (0001 → 111O) and (0001 → 2000) are found to be too small compared with experimental data. For CO2 + D2, (0001 → 1000) is also the most important process and appears to represent experimental data at room temperature. The sum of three cross sections of CO2 + H2 is always greater than that of CO2 + D2 over the temperature range of 100–2500 K.  相似文献   

4.
《Chemical physics》1987,114(3):389-397
Two reactive processes are observed: X + H2 → XH + H (R) and X + H2 → XH + H + e (RD). The angular and energy distributions of the molecular XH products are measured at collision energies varying from 5 to 10 eV center of mass. These distributions obey identical rules in the three systems: (a) XH molecules formed by both R and RD processes are scattered at the same c.m. angle, respectively 55° ±10 for ClH, 80° ±20 for BrH and 90° ±20 for IH. (b) The rovibrational energy of the XH molecules, when formed by R processes, is limited to a small amount: ⩽ 2 eV for ClH, ⩽ 1.5 eV for BrH, ⩽ 1 eV for IH, whereas when formed by RD it extends to the highest amount available from the collision energy, up to the dissociation limit. The RD process is not observed experimentally in the I/H2 system. This dynamical behaviour is fully understood in terms of non-adiabatic interaction between the two lowest [XH2] ionic surfaces, but the reason of the angular anisotropy is still not well understood.  相似文献   

5.
The temperature dependence of the removal of the vibrational energy of H2 by DCl in H2(1) + DCl(0) has been investigated over the range of 300–3000 K. The energy transfer probability of H2(1) + DCl(0) → H2(0) + DCl(1), where the vibrational energy of H2(1) is removed by both the vibrational and rotational motions of DCl(0), is found to be strongly temperature dependent and increases with temperature closely following the relation log P α T1/3. Over the temperature range it changes by two orders of magnitude. The probability of the near-resonant process H2 (1) + DCl(O) → H2(0) + DCl(2) is very close to that of the former at 300 K, but it increases only slightly as the temperature is raised to 3000 K. The sum of the probabilities of these two processes at 300 K is 3.4 × 10?5, which agrees with the experimental value of 3.95 × 10?5.  相似文献   

6.
The production of ClOO and ClO radicals following the flash photolysis of chlorine + oxygen mixtures has been studied. For the mechanism the following kinetic parameters were measured: k3K = 1.3 × 1010 l2/mol2·sec; k2/k3 = 17; and k3/?(ClOO; 250 nm) = 9.7 × 105 cm/sec. Then k3 = 5.9 × 109 l/mol·sec, k2 = 1.0 × 1011 l/mol·sec, and ?(ClOO; 250 nm) = 6.1 × 103 l/mol·cm. From limits established for the equilibrium constant K, ΔH°f (ClOO) = 94 ± 2 kJ/mol.  相似文献   

7.
The mechanism of the reactions of electronically excited SO2 with isobutane has been studied through the measurement of the initial quantum yields of product formation in 3130 Å irradiated gaseous binary mixtures of SO2 and isobutane and ternary mixtures of SO2, isobutane, C6H6 or CO2. Under low-pressure conditions (P < 10 torr) the kinetic treatment of the present data shows that only one singlet and one triplet state, presumably the 1B1 and 3B1 states, are involved in the photoreaction mechanism. The data give k2a = 8.4 × 109; SO2(1B1) + isobutane → products (2a); k5a ? k5 = 8.7 × 108 l./mol·sec; SO2(3B1) + isobutane → products (5a) SO2(3B1) + isobutane → (SO2) + isobutane (5b) k1a/k1 = 0.145 ± 0.037; SO2(1B1) + SO2 → SO2(3B1) + SO2 (1a) SO2(1B1) + SO2 → (2SO2) (1b) k2b/k2 = 0.273 ± 0.018; SO2(1B1) + isobutane → SO2(3B1) + isobutane (2b); SO2(1B1) + isobutane → (SO2) + isobutane (2c) error limits are ± 2 σ. The contribution from the excited SO2(1B1) molecules to the quantum yields of the photolyses of SO2–isobutane mixtures is not negligible. Under high-pressure conditions (P > 10 torr) the low-pressure mechanism coupled with the saturation effect on the phosphorescence lifetimes of SO2(3B1) molecules cannot alone rationalize the quantum yields. The evaluation suggests that some nonradiative intermediate state (X) is involved in the formation of “extra” triplet molecules. This ill-defined state decays largely nonradiatively to SO2 in experiments at low pressures, X → SO2 (12). In the presence of C6H6 the low-pressure data give k7 = (8.5 ± 1.8) × 1010, and the high-pressure data give k7 = (8.3 ± 0.6) × 1010 and (9.9 ± 0.9) × 1010l./mol·sec; SO2(3B1) + C6H6 → nonradiative products (7). These estimates are in good agreement with values directly measured from low-pressure lifetime studies, (8.1 ± 0.7) × 1010 and (8.8 ± 0.8) × 1010l./mol·sec.  相似文献   

8.
The H+ + LiH → Li + H reactive scattering has been studied using a quantum real wave packet method. The state‐to‐state and state‐to‐all reaction probabilities for the entitled collision have been calculated at zero total angular momentum. The probabilities for J > 0 are estimated from the J = 0 results by using J‐shifting approximation based on the Capture model. The integral cross sections and thermal rate constants are then calculated. © 2006 Wiley Periodicals, Inc. Int J Quantum Chem, 2006  相似文献   

9.
The kinetics of C2H5O2 and C2H5O2 radicals with NO have been studied at 298 K using the discharge flow technique coupled to laser induced fluorescence (LIF) and mass spectrometry analysis. The temporal profiles of C2H5O were monitored by LIF. The rate constant for C2H5O + NO → Products (2), measured in the presence of helium, has been found to be pressure dependent: k2 = (1.25±0.04) × 10?11, (1.66±0.06) × 10?11, (1.81±0.06) × 10?11 at P (He) = 0.55, 1 and 2 torr, respectively (units are cm3 molecule?1 s?1). The Lindemann-Hinshelwood analysis of these rate constant data and previous high pressure measurements indicates competition between association and disproportionation channels: C2H5O + NO + M → C2H5ONO + M (2a), C2H5O + NO → CH3CHO + HNO (2b). The following calculated average values were obtained for the low and high pressure limits of k2a and for k2b : k = (2.6±1.0) × 10?28 cm6 molecule?2 s?1, k = (3.1±0.8) × 10?11 cm3 molecule?1 s?1 and k2b ca. 8 × 10?12 cm3 molecule?1 s?1. The present value of k, obtained with He as the third body, is significantly lower than the value (2.0±1.0) × 10?27 cm6 molecule?2 s?1 recommended in air. The rate constant for the reaction C2H5O2 + NO → C2H5O + NO2 (3) has been measured at 1 torr of He from the simulation of experimental C2H5O profiles. The value obtained for k3 = (8.2±1.6) × 10?12 cm3 molecule?1 s?1 is in good agreement with previous studies using complementary methods. © 1995 John Wiley & Sons, Inc.  相似文献   

10.
The primary processes in the photolysis of water vapor at 1470 Å are due to H2O + hν(λ = 1470 Å) → H2 + O(1D), H2O + hν(λ = 1470 Å) → H + OH with the H2 yield of the first process accounting for 23% of the overall H2 production. The quantum yield of this process is estimated to be 0.08 by using O2 as a scavenger for H-atoms. Secondary reactions involving the photolytic products and added O2 are discussed.  相似文献   

11.
The reaction Ar(2P2,0) + H2O → Ar + H + OH(A2Σ+)was studied in crossed molecular beams by observing the luminescence from OH(A2Σ+). No significant dependence of the spectrum on collision energy was found over the 22–52 meV region. Spectral simulation was used to obtain the OH(A) vibrational distribution and rotational temperature, assuming a Boltzmann rotational distribution. Since predissociation is known to strongly affect the rovibrational distribution, the individual rotational state lifetimes were included in the simulation program and were used to obtain the average vibrational state lifetimes. Excellent agreement with experiment was obtained for vibrational population ratios N0/N1/N2 of 1.00/ 0.40/0.013 and a rotational temperature of 4000 K. Correction for the different average vibrational lifetimes gave formation rate ratios P0/P1/P2 of 1.00/0.49/0.25. The differences between these results and those from flowing afterglow studies on the same system are discussed. Three reaction mechanisms are considered, and the vibrational prior distributions are calculated from a simple density-of-states model. Only fair agreement with experiment is obtained. The best agreement for the mechanisms giving OH(A) in two 2-body dissociation steps is obtained by assuming 1.0 eV of internal energy remains in the second step. The OH(A) vibrational population distribution of the present work is similar to that found in the photolysis of H2O at 122 nm, where there is 1.10 eV of excess internal energy.  相似文献   

12.
The pyrolysis of 2% CH4 and 5% CH4 diluted with Ar was studied using both a single–pulse and time–resolved spectroscopic methods over the temperature range 1400–2200 K and pressure range 2.3–3.7 atm. The rate constant expressions for dissociative recombination reactions of methyl radicals, CH3 + CH3 → C2H5 + H and CH3 + CH3 → C2H4 + H2, and for C3H4 formation reaction were investigated. The simulation results required considerably lower value than that reported for CH3 + CH3 → C2H4 + H2. Propyne formation was interpreted well by reaction C2H2 + CH3P-C3H4 + H with ?? = 6.2 × 1012 exp(?17 kcal/RT) cm3 mol?1 s?1.  相似文献   

13.
We present nonadiabatic quantum dynamical calculations on the two coupled potential en-ergy surfaces (12A′ and 22A′) [J. Theor. Comput. Chem. 8, 849 (2009)] for the reaction. Initial state-resolved reaction probabilities and cross sections for the N+ND→N2+D reaction and N′+ND→N′D+N reaction for collision energies of 5 meV to 1.0 eV are determined, re-spectively. It is found that the N+ND→N2+D reaction is dominated in the N+ND reaction.In addition, we obtained the rate constants for the N+ND→N2+D reaction which demand further experimental investigations.  相似文献   

14.
Using the technique of flash photolysis-resonance fluorescence, absolute rate constants have been measured for the reaction H + O2 + M → HO2+M over a temperature range of 220–360°K. Over this temperature range, the data could be fit to an Arrhenius expression of the following form: The units for kAr are cm6/mole-s. At 300°K the relative efficiencies for the third-body gases Ar:He:H2:N2:CH4 were found to be 1.0:0.93:3.0:2.8:22. Wide variations in the photoflash intensity at several temperatures demonstrated that the reported rate constants were measured in the absence of other complex chemical processes.  相似文献   

15.
The quantum yields of the sulfur dioxide triplet (3SO2)-sensitized phosphorescence of biacetyl (Φsens) were determined in experiments with N2–SO2–Ac2 and c-C6H12–SO2–Ac2 mixtures excited at 2875 Å at 27°C. The fraction of the biacetyl triplets which reacts homogeneously by radiative or nonradiative decay reactions was determined in a series of runs at constant [SO2]/[M] and [SO2]/[Ac2] ratios but at varied total pressure. A kinetic treatment of the Φsens results and singlet sulfur dioxide (1SO2) quenching rate constant data gave the following new kinetic estimates: 1SO2 + M → (SO2–M) (1b) 1SO2 + M → 3SO2 + M (2b); for 1SO2–N2 collisions, k2b/(k1b + k2b) = 0.033 ± 0.008; for 1SO2c-C6H12 collisions, k2b/(k1b ± k2b) = 0.073 ± 0.024; previous studies have shown this ratio to be 0.095 ± 0.005 for 1SO2–SO2 collisions. It was concluded that the inter-system crossing ratio in 1SO2 induced by collision is relatively insensitive to the nature of the collision partner M. However, the individual rate constants for the collision-induced spin inversion of 1SO2 (k2b) and the total 1SO2-quenching constants (k1b + k2b) are quite sensitive to the nature of M: k2b/k2a varies from 0.10 ± 0.03 for M = N2 to 1.11 ± 0.37 for M = c-C6H12, and (k1b + k2b)/(k1a + k2a) varies from 0.29 for M = N2 to 1.44 for M = c-C6H12; k1a and k1b are the rate constants for the reactions 1SO2 - SO2 → (2SO2) (1a) and 1SO2 + SO23SO2 + SO2 (2a), respectively.  相似文献   

16.
The mechanism of formation of the electronically excited radical OH*(A2Σ+) has been studied by analyzing calculations quantitatively describing the results of shock wave experiments carried out in order to determine the moment of maximum OH* radiation at temperatures T < 1500 K and pressures P ≤ 2 atm in the H2 + O2 mixtures diluted by argon when the vibrational nonequilibrium is a factor determining the mechanism and rate of the overall process. In kinetic calculations, the vibrational nonequilibrium of the initial H2 and O2 components, the HO2, OH(X2Π), O2*(1Δ) intermediates, and the reaction product H2O were taken into account. The analysis showed that under these conditions the main contribution to the overall process of OH* formation is caused by the reactions OH + Ar → OH* + Ar, H2 + HO2 → OH* + H2O, H2 + O*(1D) → OH* + H, HO2 + O → OH* + O2 and H + H2O → OH* + H2, which occur in the vibrational nonequilibrium mode (their activation barrier is overcome due to the vibrational excitation of reactants), and by H + O3 → OH* + O2 and H + H2O2 → OH* + H2O, which are reverse to the reactions of chemical quenching.  相似文献   

17.
The reaction of OH with NOCl has been studied using the discharge flow reaction-EPR technique. The absolute rate constant is k1 = (4.3±0.4)× 10?13 cm3 molecule?1 s?1 at 298 K. A mass spectrometric investigation of the products shows that this reaction occurs via two primary steps, OH + NOCl → NO + ClOH(1a) and OH + NOCl → HONO + Cl (1b) with k1a =k1b.  相似文献   

18.
Weakly bound molecular complexes with more than one well-defined structures provide us with an unique opportunity to investigate dynamic processes induced by intermolecular interactions with specific orientations. The relative orientation of the two interacting molecules or atoms is defined by the complex structure. The effect of the orientation in the spin changing collisions glyoxal(S1)+Ar → glyoxal(T1)+Ar and acetylene (S1)+Ar → acetylene(T)+Ar have been studied by measuring the intersystem crossing (ISC) rates of the glyoxal(S1)·Ar and acetylene(S1)·Ar complexes with different isomeric structures. Results show that there is a strong orientation dependence in the ISC of glyoxal(S1) induced by interaction with the Ar atom: the Ar atom positioned in the molecular plane is much more effective than in the out-of-plane position in inducing the S1 → T1 transition of glyoxal. On the other hand, studies of acetylene(S1)·Ar complexes indicate that the Ar-induced ISC rates are nearly identical for the in-plane and out-of-plane positions. Orientation dependence in the collision induced vibrational relaxation process C2H2(S1,v i )+Ar → C2H2(S1,v f <v i )+Ar is also studied by measuring the vibrational predissociation rates of the acetylene(S1)·Ar complex isomers. The results indicate that collisions of C2H2(S1,v 3=3, 4) with Ar at two orthogonal orientations are equally effective in causing vibrational relaxation of C2H2.  相似文献   

19.
The coupling between VV and VT transitions in H2(2) + D2(0) → H2(1) + D2(1) and D2(2) +H2(0) → D2(1) + H2(1) at high collision energies is investigated by use of the solution of the Schrodinger equation of motion. The coupling is sufficiently important that the energy exchange processes cannot be described by the VV mechanism alone.  相似文献   

20.
Measurements of the NO-catalyzed dissociation of I2 in Ar in incident shock waves were carried out in the temperature range of 700°-1520°K and at total concentrations of 5 × 10?6-6 × 10?5 mol/cm3, using ultraviolet-visible absorption techniques to monitor the disappearance of I2. It was shown that the main reaction responsible for the disappearance under these conditions is I2 + NO → INO + I, for which a rate coefficient of (2.9 ± 0.5) × 1013 exp[-(18.0 ± 0.6 kcal/mol)/RT] cm2/mol·sec was determined. The INO formed dissociates rapidly in a subsequent reaction. The reaction, therefore, constitutes a “chemical model” for a “thermal collisional release mechanism.” Preliminary measurements of the rate coefficient for I2 + NO2 → INO2 + I are also presented. Combined with information on the reverse reactions obtained in earlier room temperature experiments, these results lead to accurate values of ΔH°f for INO and INO2 equal to 29.7 ± 0.5 and 15.9 ± 1 kcal/mol, respectively.  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