Quasi-classical trajectory study of the stereodynamics of a Ne+H + 2 NeH+ +H reaction

Similar documents
Stereodynamics of the O( 3 P) with H 2 (D 2 ) (ν = 0, j = 0) reaction

Isotope effect on the stereodynamics for the collision reaction H+LiF(v = 0, j = 0) HF+Li

Effect of isotope substitution on the stereodynamics for O+H(D)Br OH(D)+Br reactions. 1 Introduction

Equivalence between Symmetric and Antisymmetric Stretching Modes of NH 3 in

Photodetachment of H in an electric field between two parallel interfaces

Chemical Kinetics and Reaction Dynamics Group UR UB. List of publications. Updated up to 2007.

Calculations of the binding energies of weakly bound He He H, He He H and He H H molecules

Resonances in Chemical Reactions : Theory and Experiment. Toshiyuki Takayanagi Saitama University Department of Chemistry

1 Molecular collisions

Procedia Computer Science

The rotating Morse potential energy eigenvalues solved by using the analytical transfer matrix method

Electron detachment process in collisions of negative hydrogen ions with hydrogen molecules

Transport properties through double-magnetic-barrier structures in graphene

A Combined Experimental-Theoretical Study of the OH + CO H + CO 2. Reaction Dynamics

Sunlight loss for femtosecond microstructured silicon with two impurity bands

Brazilian Journal of Physics, vol. 36, no. 3A, September,

CH Stretching Excitation Promotes its Cleavage in. Collision Energies

Atom-molecule molecule collisions in spin-polarized polarized alkalis: potential energy surfaces and quantum dynamics

Shape Coexistence and Band Termination in Doubly Magic Nucleus 40 Ca

Coriolis Coupling Influence on the H+LiH Reaction

New Feedback Control Model in the Lattice Hydrodynamic Model Considering the Historic Optimal Velocity Difference Effect

D. De Fazio, T. V. Tscherbul 2, S. Cavalli 3, and V. Aquilanti 3

Hongwei Song and Hua Guo * Department of Chemistry and Chemical Biology, University of New Mexico, Albuquerque, New Mexico, 87131, USA

Used for MS Short Course at Tsinghua by R. Graham Cooks, Hao Chen, Zheng Ouyang, Andy Tao, Yu Xia and Lingjun Li

Structure stability and magnetic properties of Os n B(n = 11 20) clusters

Isotopic effect of Cl + 2 rovibronic spectra in the A X system

c re l 4. Controlling reagents and characterizing products Control of reagent orientation

Density Functional Theory Study on Mechanism of Forming Spiro-Geheterocyclic Ring Compound from Me 2 Ge=Ge: and Acetaldehyde

Double-distance propagation of Gaussian beams passing through a tilted cat-eye optical lens in a turbulent atmosphere

The near-infrared spectra and distribution of excited states of electrodeless discharge rubidium vapour lamps

Full dimensional quantum scattering study of the H 2 + CN reaction #

Supporting Information. I. A refined two-state diabatic potential matrix

Spin-orbit effect in the energy pooling reaction

A Reactant-Coordinate-Based Wave Packet Method for Full-dimensional State-to-State

Effects of aligning pulse duration on the degree and the slope of nitrogen field-free alignment

Cl HD v 1; J 1,2 reaction dynamics: Comparison between theory and experiment

Generalized projective synchronization between two chaotic gyros with nonlinear damping

Influence of Isotope Effects on Product Polarizations of N( 2 D)+D 2, N( 2 D)+H 2 and N( 2 D)+HD Reactive Systems

Photodetachment of H Near a Dielectric Surface

Effects of Vibrational Excitation on the F + H2O HF + OH Reaction: Dissociative Photodetachment of Overtone Excited [F H OH]ˉ

A six dimensional quantum study for atom triatom reactions: The H H 2 O H 2 OH reaction

3D Time-Dependent Wave-Packet Approach in Hyperspherical Coordinates for the H+O 2 Reaction on the CHIPR and DMBE IV Potential Energy Surfaces

Wavepacket Correlation Function Approach for Nonadiabatic Reactions: Quasi-Jahn-Teller Model

Resonant enhanced electron impact dissociation of molecules

Plot the interatomic distances as a function of time and characterize the reactants and products through the plot. w

Fragmentation of Na 3 clusters following He impact: Theoretical analysis of fragmentation mechanisms

Candidate multiple chiral doublet bands in A 100 mass region

Calculation of Molecular Constants for the of the NeH + and KrH + Ions

Chemical reactions occur when one reactant collides with another

Effects of Isospin on Pre-scission Particle Multiplicity of Heavy Systems and Its Excitation Energy Dependence

Semi-Relativistic Reflection and Transmission Coefficients for Two Spinless Particles Separated by a Rectangular-Shaped Potential Barrier

A comparative study of the Si+O2 SiO+O reaction dynamics from quasiclassical trajectory and statistical based methods

Diagrammatic Representation of Electronic Correlations in Photoionization Process: Application to Scandium

Analysis of second-harmonic generation microscopy under refractive index mismatch

1292 Wang Shu-Xia et al Vol. 12 those in YBa 2 Cu 3 O 7 x. But as far as the subtle characters are concerned, ErBa 2 Cu 3 O 7 x is more similar to YBa

Scheme for Asymmetric and Deterministic Controlled Bidirectional Joint Remote State Preparation

The calculation of energy gaps in small single-walled carbon nanotubes within a symmetry-adapted tight-binding model

Combined Influence of Off-diagonal System Tensors and Potential Valley Returning of Optimal Path

Stopping power for MeV 12 C ions in solids

Roaming and Spectroscopy

Vibrationally elastic and inelastic scattering of electrons by hydrogen sulphide molecules

Photoelectron Spectroscopy of the Hydroxymethoxide Anion, H 2 C(OH)O

Absolute Integral and Differential Cross Sections for the Reactive Scattering of H - + D 2 and D - + H 2

Intermediate Energy Pion- 20 Ne Elastic Scattering in the α+ 16 O Model of 20 Ne

Chaos suppression of uncertain gyros in a given finite time

Supplementary Information

Quantum chemistry and vibrational spectra

No. 2 lectronic state and potential energy function for UH where ρ = r r e, r being the interatomic distance and r e its equilibrium value. How

Radiation energy flux of Dirac field of static spherically symmetric black holes

Vibrationally Mediated Bond Selective Dissociative Chemisorption of HOD on Cu(111) Supporting Information

Throwing Light on Reaction Dynamics: H + HBr

Projective synchronization of a complex network with different fractional order chaos nodes

State-to-State Kinetics of Molecular and Atomic Hydrogen Plasmas

1 Molecular collisions

Azimuthal Asymmetries of the Drell-Yan Process in pa Collisions

Assignment: Read Atkins, Chapter 27 sections 7 and 8 or McQuarrie and Simon, Chapter 30 sections 7 and 10, before coming to lab on Monday

Non-Equilibrium Reaction Rates in Hydrogen Combustion

Effects of Atomic Coherence and Injected Classical Field on Chaotic Dynamics of Non-degenerate Cascade Two-Photon Lasers

Generalized projective synchronization of a class of chaotic (hyperchaotic) systems with uncertain parameters

Seven dimensional quantum dynamics study of the H 2 +NH 2 \H+NH 3 reaction

A study of φ-meson spin alignment with the AMPT model

Time-delay feedback control in a delayed dynamical chaos system and its applications

Scattering of Xe from graphite. Abstract

5.80 Small-Molecule Spectroscopy and Dynamics

Quantum-dynamical Characterization of Reactive Transition States

Electron-Vibrational Energy Exchange in Nitrogen-Containing Plasma: a Comparison Between an Analytical Approach and a Kinetic Model

Centrifugal Barrier Effects and Determination of Interaction Radius

Backstepping synchronization of uncertain chaotic systems by a single driving variable

Calculation of the State-to-state S-Matrix for Tetra-atomic Reactions with. Albuquerque, New Mexico 87131, USA

Construction of a New Fractional Chaotic System and Generalized Synchronization

Optical time-domain differentiation based on intensive differential group delay

Surface-hopping trajectories for OH(A2Σ+) + Kr: Extension to the 1A state

Electron Transport Behavior in a Mirror Magnetic Field and a Non-uniform Electric Field

MULTIPACTOR ON A DIELECTRIC SURFACE WITH LONGITUDINAL RF ELECTRIC FIELD ACTION

Raman spectroscopy at the edges of multilayer graphene

arxiv:cond-mat/ v1 [cond-mat.other] 5 Jun 2004

On Factorization of Coupled Channel Scattering S Matrices

Antimagnetic rotation in 108,110 In with tilted axis cranking relativistic mean-field approach *

Crossed-beams and Theoretical Studies of Hyperthermal Reactions of O(3P) with HCl

Absorption-Amplification Response with or Without Spontaneously Generated Coherence in a Coherent Four-Level Atomic Medium

Transcription:

Quasi-classical trajectory study of the stereodynamics of a Ne+H + 2 NeH+ +H reaction Ge Mei-Hua( ) and Zheng Yu-Jun( ) School of Physics, Shandong University, Jinan 250100, China (Received 19 February 2011; revised manuscript received 21 March 2011) We have carried out a quasi-classical trajectory calculation for the reaction of Ne + H + 2 (ν = 0, j = 1) NeH+ + H on the ground state (1 2 A ) using the LZHH potential energy surface constructed by Lü et al. [Lü S J, Zhang P Y, Han K L and He G Z 2010 J. Chem. Phys. 132 014303]. Differential cross sections at many collision energies indicate that the reaction is dominated by forward-scattering. In addition, the NeH + product shows rotationally hot and vibrationally cold distributions. Stereodynamical results indicate that the products are strongly polarized in the direction perpendicular to the scattering plane and that the products rotate mainly in planes parallel to the scattering plane. Keywords: quasi-classical trajectory, stereodynamics, product vibrational distribution, product rotational distribution PACS: 34.50.Lf, 82.20.Kh, 82.20.Pm DOI: 10.1088/1674-1056/20/8/083401 1. Introduction Recently, atom molecule ion reactions and the related systems have attracted much attention [1,2] since they can play an important role in many interesting cases including interstellar processes, electric discharges and planetary ionospheres. Rare gashydrogen systems have continued to provide an important insight into chemical reaction dynamics. So Ne+H + 2 NeH + +H reaction has been the subject of numerous experimental [3 7] and theoretical [4,8 19] studies. In order to obtain accurate potential energy surfaces (PESs) for the title reaction, many efforts have been made. [8,15 19] There are also many dynamical calculations in order to calculate reaction probabilities, [8 14,17] integral cross sections (ICSs) [4,9 13,16] and product rotational distributions (PRDs), [11] and so on. Almost all these dynamical calculations [9 14] are based on the same analytical PES in Ref. [15], while the initial rotational state ( j = 0) and collision energies for these calculations are different from those in Refs. [4] and [8]. Recently, Lü et al. [8] presented a new LZHH PES for the ground state (1 2 A ) of the title reaction from accurate ab initio data and carried out an exact quantum scattering study as well. The well in their PES [8] is 0.026 ev deeper (when the Ne H H angle is 180 ) and the energy barrier is 0.075 ev (when the Ne H H angle is 90 ) lower than the corresponding ones of Ref. [15]. Their calculated cross sections including Coriolis coupling (CC) fit well with the experimental ones. [4] However, a relatively large difference exists between the corrected quasi-classical trajectory (QCT) cross sections of Zhang et al. [4] and those of the quantum mechanical (QM) calculation. [8] There is a discrepancy between their CC ICSs and the results of previous studies. [4,9 13,16] Lü et al. [8] assumed that the difference was ascribed to the use of different PESs. According to selection rules for homonuclear molecules, transition from excitation states with odd (or even) rotational quantum numbers to the rovibrational state of ν = 0, and j = 0 (or j = 1) is forbidden. So only transition between two odd (or even) rotational states is allowed. Lü et al. [8] and Zhang et al. [4] considered the initial rotational state as j = 1 in the QM calculation and experimental study, respectively. So one can find that it is necessary to study the dynamical characteristics of the ν = 0 and j = 1 state. As exhibited in Ref. [20], the characteristic of a potential energy surface could be evaluated in classical trajectory calculations. The QCT method constitutes a good approximation to the scattering dynamical properties and the qualification of a dynam- Project supported by the National Natural Science Foundation of China (Grant No. 21073110) and the Independent Innovation Foundation of Shandong University of China (Grant No. 10000059614011). Corresponding author. E-mail: mhge@sdu.edu.cn 2011 Chinese Physical Society and IOP Publishing Ltd http://www.iop.org/journals/cpb http://cpb.iphy.ac.cn 083401-1

ical feature as a quantum effect in a given reactive system requires the performance of QCT calculations on the same PES. So we present a quasi-classical calculation of Ne + H + 2 NeH+ + H on the 1 2 A PES given by Lü et al. [8] with the initial state of H + 2 at ν = 0 and j = 1 so as to testify whether the disparity between the results of Lü et al. and those of former researchers lies in the difference between PESs. The total cross sections are calculated to compare with CC results of Lü et al., [8] the experimental ICSs, [4] the fitted experimental results through the line-of-centre collision (LOC) model [4] and those of corrected QCT [4] as well. With the development of laser techniques, the accurate measurement of the reagent/product orientation and alignment in molecular reactions has become available. To rationalize experimental results, one needs the assistance of computational work. Generally, theoretical exploration on the stereodynamics falls into two categories: one is for those of reagent k j, j k k vector correlations and the other is for those of product k j, k k j vector correlations, where k/k and j/j denote the relative velocity vectors of reagent/product and the reagent/product rotational angular momentum in the centre of mass (CM) frame, respectively. Only limited theoretical calculations, using either quasi-classical trajectory methods, quantum scattering, wave packet propagation techniques or simple models, [21 34] have been made of angular momentum polarization. As far as we know, up to now few stereodynamical characteristics of the title reaction have been studied, so we provide some stereodynamical information about the product k j, k k j vector correlations. Differential cross sections (DCSs) as a function of both the scattering angle and collision energy are calculated, which exhibits the sensitivity of DCSs to the scattering angle and collision energy, and the tendency of scattering direction of NeH +. There is also an exhibition of product vibrational and rotational distributions over the considered collision energy interval. In addition, he polar angular distribution, P (θ r ), the distribution of the dihedral angle denoting k k j correlation, P (φ r ) and the angular distribution of product rotational vectors in the form of polar plots P (θ r, φ r ) are calculated to validate more information about the angular momentum polarization and orientation. 2. Computational aspects Our calculation is performed on 1 2 A PES presented by Lü et al. [8] As is shown in Fig. 1, there is a deep well in the collinear geometry and its depth decreases with Ne H H angle decreasing, and at some point the well turns into barrier. The barrier height is 0.425 ev in perpendicular geometry. The ground state surface has an overall minimum in collinear geometry and the well is about 0.536 ev deep with respect to the reaction asymptote. For the details of the PES, readers may refer to Ref. [8]. Based on classical Fig. 1. Contour plots of the PES for the four different Ne H H angles in internal coordinates (180, 120, 90, and 60 ). The dotted line denotes 0 ev relative to the energy of the reactant Ne+ H + 2. 083401-2

trajectory method, in the QCT method some important quantum effects (using the vibrational and rotational quantum numbers to describe the reagent/product state) are taken into consideration, and the QCT method has been used to study the reaction dynamics of many systems. The calculation method we used is the same as the one described in Refs. [21] [30]. In QCT calculation, there is uncertainty in determining the optimized impact parameter b max. The associated uncertainties with the QCT total cross section can be calculated by σ = [(N tot N r )/(N tot N r )] 1/2 σ. Therefore, we estimate the error margin for the present QCT ICS to be less than ±6.7%. 3. Results and discussion The classical Hamilton equation is solved in three dimensions. We carry out the calculation with the initial vibrational and rotational numbers ν = 0 and j = 1, respectively. The values of translational energy E col range from 0.4 ev to 1.5 ev with the energy gap 0.1 ev. The time integral step size is chosen as 0.1 fs and 50000 trajectories are used in the computation. In Fig. 2, our QCT ICSs and the associated errors are compared with previous results, [4,8] including quantum dynamical (CC) results, the experimental measurements, the deconvoluted LOC results (which are fitted to the ICS curve according to their experimental results by using the modified LOC threshold function) and corrected QCT results (which refers to proton transfer cross sections corrected for NeH+ products with vibrational energies below the zero point energy). Peaking at E col = 1.0 ev, our QCT ICSs display an increasing-then-decreasing trend over the whole collision energy range. For the CC and the experimental results, the phenomena are similar. Our calculated threshold is a bit smaller than that of CC computation and the deconvoluted one, owning to a well-known handicap of classical mechanics as zero point energy leakage. [35,36] Over the considered energy range, QCT ICSs are smaller than the experimental ones. The same situation also occurred in Ref. [13], which stated that the unusually large QCT cross sections at the threshold are mainly because of the inability of classical mechanics to handle correctly the zero-point energy of the products and the smaller QCT cross sections at higher energies are due to a combination of tunneling processes and the existence of the resonances. So we ascribe the discrepancy to the neglect of quantum effect in the QCT calculation. Except for the case in Ref. [16], different initial ro-vibrational states are chosen for the ICS calculation, moreover, the authors used a different PES in Ref. [16], which might result in the discrepancy between the CC results of Lü et al. and previous results. The main difference between the PESs used in the present paper and that in Ref. [14] is the well depth and the energy barrier. The former PES [8] has a deeper well depth for collinear geometry and a lower energy barrier when the Ne H H angle is 90. The differences are 0.06 ev and 0.115 ev, respectively. In a word, the magnitude and profile of present QCT ICSs accord well with the CC and experimental results compared with previous results. In addition, the smaller QCT ICSs at higher collision energies (E col > 1.0 ev) perhaps result from the large number of resonances (which could be observed in reaction probabilities of Lü et al. [8] ) and tunneling processes (which might be ascribed to the deep well at large Ne H H angles). Fig. 2. Comparison between present ICSs (filled circles) and previous ones (CC of Lü et al. (solid line); experimental measurements of Zhang et al. (open circles); the deconvoluted LOC results (dashed line); and corrected QCT cross sections (open triangles). And the error bars (2.6% 6.7%) for the QCT ICSs represent the maximum absolute errors (1 Å=0.1 nm). DCSs as a function of both the scattering angle (the angle between the reagent relative velocity k and the product velocity k ) and collision energy E col are given in Fig. 3 so as to explain the tendency of scattering direction of the product. For each of the considered energies there exists a maximum when the scattering angle θ t is 36.9, which suggests that the forward-reaction is dominant. The forward type distribution is consistent to a certain extent with 083401-3

the experimental results in Ref. [7], which were obtained at E col = 0.87 ev (namely within the range of our translational energies). Due to the fact that the ro-vibrational states are not distinguished from the reagent states, a linear combination of all H + 2 states (probably a Franck Condon distribution) is applied to the experimental results. Their experimental measurements [7] result in a strong forward distribution. The agreement between our calculations and the experimental results is due to two reasons: (i) vibrational excitation greatly enhances reactivity, [8,11,12,17] and according to the results of Ref. [12], the DCSs of ν = 1, 2 in their studied energy interval (which is within the energy range of our calculation) also demonstrate strong forward scattering; (ii) as expected, the contribution of all higher vibrational states will also be in the forward direction, as they all lead to exothermic reactions. From Fig. 3, one can observe that DCS is sensitive to the collision energy, especially for small scattering angle. Also DCS displays an increasing-then-decreasing trend over the whole collision energy range, and the peak appears at a collision energy of 1.0 ev, which is in accordance with the ICS result. main interesting point is that if the increased collision energy is not larger than the energy gap between two vibrational states of the reagent, the number of the vibrational channel to be open would not increase, but it could have influence on the distribution function for a defined v, especially for small collision energies. For example, when the collision energy varies from 0.4 ev to 0.6 ev, the number of the open vibrational channel remains unchanged (v = 0) with a distribution ratio 1:5.8:12.7 (from small to large collision energy). Figure 4(b) exhibits the PRDs for the collision energies of 0.4 ev 1.5 ev. Overall PRDs demonstrate a unimodal distribution at almost all the collision energies, peaking at an intermediate value of the open product rotational quantum numbers. More scanned product rotational levels are open when the collision energy is increased. Another interesting feature exhibited in Fig. 4(b) is the dependence of PRD on collision energy. It is not the same for each collision energy. The most striking effect can be observed when the collision energy goes from 0.4 ev to 0.6 ev. We find that high states become more populated as the collision energy increases, yielding a distribution peaked at j = 1 at 0.4 ev, j = 2 at 0.5 ev, and j = 3 at 0.6 ev, which implies that the peak shifts towards larger rotational quantum numbers with the increase of collision energy. Fig. 3. DCSs as a function of both the scattering angle θ t and collision energy E col for the Ne + H + 2 (ν = 0, j = 1) NeH + +H reaction. The QCT method allows for a complete state-tostate study of the reaction dynamics, so we calculate the product vibrational and rotational distributions of NeH + as shown in Fig. 4. From Fig. 4(a), we find that only four product vibrational levels at the highest energies scanned here are open, and the product vibrational distributions (PVDs) all peak at v = 0 over the whole energy range. In general, the lowest vibrational state is favoured and the distribution appears to be cold with a probability of zero above v = 3. The Fig. 4. Product ro-vibrational distributions for the Ne + H + 2 (ν = 0, j = 1) NeH+ (v, j ) + H reaction: (a) PVDs, (b) PRDs. 083401-4

The vibrational and the rotational excitations of the product exhibit different distribution functions. Through the comparison between Figs. 4(a) and 4(b), we can conclude that the NeH + product is rotationally hot and vibrationally cold. Owing to the large number resonances in reaction probabilities of Lü et al., [8] the title reaction is dominated by the formation of a complex. As shown in Fig. 1, the deep well in the PES accounts for such a complex-forming scheme. The complex-forming mechanism usually favours the molecule rotating in various directions. As j increases the rotational angular momentum becomes increasingly polarized. It might explain the wider range of product rotational population, and this assumption can be further testified through the depictions of P (θ r ) and P (φ r ) distributions in Figs. 5 and 6. P (φ r ) distributions are asymmetric with respect to k k plane (the scattering xz plane) with two peaks appearing nearly at φ r = 90 and φ r = 270, which demonstrates that j is preferentially aligned along y axis. The larger peak existing nearly at φ r = 90 or φ r = 270 indicates a stronger product orientation along the positive or negative direction of y axis. It is interesting that the P (φ r ) distributions are asymmetric although the title reaction is symmetric about the initial relative velocity vector. Orientation of the product maybe results from the repulsive energy between the two atoms of H + 2, which leads to the violation of the symmetry. Through comparison, we find that the orientations along the positive and the negative y axis coexist at the considered energy range, and the dominant +y axis or y axis orientation appears at medium collision energies. Moreover, the direction and the degree of the alignment/orientation change with collision energy. Fig. 5. The distribution of P (θ r) as a function of both the polar angle θ r and collision energy E col for the reaction of Ne + H + 2 (ν = 0, j = 1) NeH+ + H. To obtain a graphical representation of the polarization we plot the P (θ r ) distribution as a function of both the polar angle θ r and collision energy E col in Fig. 5. Obviously, there is a demonstration of symmetric distributions peaked at θ r = 90 over the whole energy range, implying that the product angular momentum tends to align in the direction perpendicular to k. Since higher and narrower distribution indicates stronger product alignment, one can observe that the alignment degree is much larger at low collision energies (e.g. 0.4 ev and 0.5 ev) than that at higher collision energies. There is a non-monotonic relationship between the alignment degree and the collision energy. The structure of PES (with a deep well and a late barrier) might give a reason of such a phenomenon. The P (φ r ) distributions are given in Fig. 6, which can provide some stereodynamical information about both the alignment and the orientation characteristics of the product. The φ r denotes the azimuthal angle of the final rotational angular momentum j. All Fig. 6. The P (φ r) distributions as a function of the dihedral angle φ r at the collision energies of (a) 0.4, 0.5, 0.6 ev; (b) 0.7, 0.8, 0.9 ev; (c) 1.0, 1.1, 1.2 ev; and (d) 1.3, 1.4, 1.5 ev (from inner to outer) for Ne + H + 2 (ν = 0, j = 1) NeH + + H reaction. To further examine the tendencies, we also plot the angular momentum polarization in the form of polar plots of θ r and φ r averaged over all scattering angles as shown in Fig. 7. The features of P (θ r, φ r ) distributions with peaks and valleys are in good accordance with those of the P (θ r ) and P (φ r ) distributions of the products. Also the depictions in Fig. 7 indicate that the products are strongly polarized in the direction perpendicular to the scattering plane (x z plane is the scattering plane containing the initial and the final relative velocity vectors, k and k ) and the products rotate mainly in planes parallel to the scattering plane. 083401-5

Fig. 7. Joint P (θ r, φ r) distributions as a function of both polar angle θ r and φ r for the Ne + H + 2 j = 1) NeH + + H reaction. (ν = 0, 4. Conclusion QCT calculation is performed to investigate dynamical information about the title reaction at the initial ro-vibrational state ν = 0, and j = 1. Compared with previous results, the calculated cross sections are close to the accurate QM and experimental ones, which proves the validity of QCT calculation. According to our QCT results, the disagreement between the CC ICSs of Lü et al. and former results is probably due to employing different PESs and different initial states. DCSs as a function of both the scattering angle and translational energy are given, which manifests that the forward-scattering is dominant and DCS is not a monotonic function of collision energy. Also, the product vibrational and rotational distributions are computed. The product vibrational state population is rather cold (the lowest v = 0 vibrational state of NeH + is favoured) and PRDs are highly selective (medium rotational numbers are favoured). Our results show a high rotational excitation in the product state distribution. Over the range of the considered collision energies, according to the vector properties of the title reaction, distribution P (θ r ) is symmetric with respect to 90, and the dihedral angle distribution P (φ r ) manifests j preferentially aligned along the y axis. Also the distributions of P (θ r, φ r ) showing the products rotate mainly 083401-6

in planes parallel to the scattering plane are in good accordance with the distributions P (θ r ) and P (φ r ). Acknowledgement The authors are very grateful to Prof. Han Ke-Li for providing the computational codes and the potential energy surface. References [1] Chu T S and Han K L 2008 Phys. Chem. Chem. Phys. 10 2431 [2] Chu T S, Zhang Y and Han K L 2006 Int. Rev. Phys. Chem. 25 201 [3] Dressler R A, Chiu Y, Levandier D J, Tang X N, Hou Y, Chang C, Houchins C, Xu H and Ng C Y 2006 J. Chem. Phys. 125 132306 [4] Zhang T, Qian X M, Tang X N, Ng C Y, Chiu Y, Levandier D J, Miller J S and Dressler R A 2003 J. Chem. Phys. 119 10175, and references therein [5] Herman Z and Koyano I 1987 J. Chem. Soc. Faraday Trans. 83 127 [6] Van Pijkeren D, Boltjes E, Eck J V and Niehaus A 1984 Chem. Phys. 91 293 [7] Bilotta R M and Farrar J M 1981 J. Chem. Phys. 75 1776 [8] Lü S J, Zhang P Y, Han K L and He G Z 2010 J. Chem. Phys. 132 014303 [9] Mayneris J, Sierra J D and González M 2008 J. Chem. Phys. 128 194307 [10] Huarte-Larrañaga F, Giménez X, Lucas J M, Aguilar A and Launay J M 2000 J. Phys. Chem. A 104 10227 [11] Huarte-Larrañaga F, Giménez X, Lucas J M, Aguilar A and Launay J M 1999 Phys. Chem. Chem. Phys. 1 1125 [12] Gilibert M, Giménez X, Huarte-Larrañaga F, González M, Aguilar A, Last I and Baer M 1999 J. Chem. Phys. 110 6278 [13] Gilibert M, Blasco R M, González M, Giménez X, Aguilar A, Last I and Baer M 1997 J. Phys. Chem. A 101 6821 [14] Kress J D, Walker R B, Hayes E F and Pendergast P 1994 J. Chem. Phys. 100 2728 [15] Pendergast P, Heck J M, Hayes E F and Jaquet R 1993 J. Chem. Phys. 98 4543, and references therein [16] Urban J, Klimo V, Staemmler V and Jaquet R 1991 Z. Phys. D: At. Mol. Clusters 21 329 [17] Urban J, Jaquet R and Staemmler V 1990 Int. J. Quantum Chem. 38 339 [18] Hayes E F, Siu A K Q, Chapman F M and Matcha R L 1976 J. Chem. Phys. 65 1901 [19] Kuntz P J and Roach A C 1972 J. Chem. Soc. Faraday Trans. 68 259 [20] Polanyi J C and Wong W H 1969 J. Chem. Phys. 51 1439 [21] Zhang X and Han K L 2006 Int. Quantum Chem. 106 1815 [22] Li R J, Han K L, Li F E, Lu R C, He G Z and Lou N Q 1994 Chem. Phys. Lett. 220 281 [23] Han K L, He G Z and Lou N Q 1996 J. Chem. Phys. 105 8699, and references therein [24] Han K L, Zheng X G, Sun B F, He G Z and Zhang R Q 1991 Chem. Phys. Lett. 181 474 [25] Liu S L and Shi Y 2011 Chin. Phys. B 20 013404 [26] Han B R, Zong F J, Wang C L, Ma W Y and Zhou J H 2010 Chem. Phys. 374 94 [27] Ge M H and Zheng Y J 2011 Theor. Chem. Acc. 129 173 [28] Kong H, Liu X G, Xu W W, Liang J J and Zhang Q G 2009 Acta Phys. Sin. 58 6926 (in Chinese) [29] Xu Y, Zhao J, Yue D G, Zheng X Y and Meng Q T 2009 Chin. Phys. B 18 5308 [30] Zhao J, Xu Y and Meng Q T 2010 Chin. Phys. B 19 063403 [31] Hartree W S, Simons J P and Gonzalez-Urena A 1990 J. Chem. Soc. Faraday Trans. 86 17 [32] Prisant M G, Rettner C T and Zare R N 1984 J. Chem. Phys. 81 2699 [33] Chu T S 2010 J. Comput. Chem. 31 1385 [34] Li B and Han K L 2009 J. Phys. Chem. A 113 10189 [35] Ju L P, Han K L and Zhang J Z H 2009 J. Comput. Chem. 30 305 [36] Varandas A J C 2000 Int. Rev. Phys. Chem. 19 199 083401-7