Quantum Dynamics Study of the Potential Energy Minima Effect on

Mar 27, 2017 - Quantum Dynamics Study of the Potential Energy Minima Effect on Energy Efficiency for the F– + CH3Cl → FCH3 + Cl– Reaction .... a...
2 downloads 7 Views 1MB Size
Subscriber access provided by UB + Fachbibliothek Chemie | (FU-Bibliothekssystem)

Article

Quantum Dynamics Study of Potential Energy Minima Effect on Energy Efficiency for the F + CHCl # FCH + Cl Reaction –

3

3



Yida Li, Yuping Wang, and Dunyou Wang J. Phys. Chem. A, Just Accepted Manuscript • DOI: 10.1021/acs.jpca.7b01547 • Publication Date (Web): 27 Mar 2017 Downloaded from http://pubs.acs.org on March 30, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry A is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Quantum Dynamics Study of Potential Energy –

Minima Effect on Energy Efficiency for the F + –

CH3Cl → FCH3 + Cl Reaction Yida Li, Yuping Wang, and Dunyou Wang* College of Physics and Electronics, Shandong Normal University, Jinan 250014, Shandong, China

* Author to whom correspondence should be addressed. Email: [email protected]

1

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 29

ABSTRACT The Polanyi rules on the energy efficiency on reactivity are summarized solely from the locations of barriers on the potential energy surfaces. Here our quantum dynamics –



study for the F + CH3Cl → FCH3 + Cl reaction shows that the two potential energy minima in the entrance channel on the potential energy surface play an essential role in energy efficiency on reactivity. The reactivity of this reaction is dominated by the low collision energies where two distinctive reaction mechanisms involve the two minima in the entrance channel. Overall, the Cl-CH3 stretching motion and C-H3 umbrella motion both are more efficient than the translational motion in promoting this reaction. Although this reaction has a negative energy barrier, our study shows that it is the minima in the entrance channel, together with the energy barrier relative to these minima, that determine the energy efficacy on reactivity.

2

ACS Paragon Plus Environment

Page 3 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

I. INTRODUCTION The Polanyi rules,1 derived from the atom-diatom reactions, state that for late barrier on potential energy surface(PES), the vibrational energy is more effective than translational energy in surmounting the energy barrier into products, and the reverse is true for an early barrier reaction. Thus the Polanyi rules are entirely based on the location of the transition state on the PESs for the tri-atomic reaction systems. In recent years, both experimental and theoretical studies2-28 have been shown that there does not exist a unified rule on the energy efficacy regarding the locations of transition states on more than three-atom reaction systems. The PESs of the poly-atomic reaction systems are more complex usually with potential energy minima in both the reactant and product channels. For example, for the F + H2O reaction23, its PES has an early barrier with two main van der Waals(vdW) wells(one in the reactant channel, the other product channel). There is also an additional vdW complex in the product channel not directly connected to the transition state. For this reaction, the reactant H2O vibrational energy is more effective than translational energy in promoting the reaction,7 which contradicts to the Polanyi rules. For the O + CH4/CD4/CHD3 reaction with a slightly late barrier, where the PES has two vdW wells in the entrance channel and one in the product channel,24 studies13-16 on the reactions show that the translational energy is more effective than all the vibrational motions in surmounting the slightly late barrier. The Polanyi rules cannot be applied to this system either. Then for the Cl + CHD3 reaction system with a later barrier,25 it also has two vdW wells in the entrance channel and one in the 3

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 29

product channel, both experimental5 and theoretical study4 confirmed that overall the C-H vibrational motion is more effective than the translational motion for promoting reactivity. Hence, for this reaction, the Polanyi rules are still upheld. Therefore, if one only considers the locations of transition states on the PESs, the Polanyi rules cannot be simply extended to the poly-atomic reaction systems. Since the poly-atomic reactions usually have energy minima on their PESs, which makes one wonder whether the energy minima, especially the entrance channel wells before the transition state, might also play a role that cannot be neglected on energy efficacy on reactivity. There have been extensive studies on Nucleophilic Substitution (SN2) reactions due to their great importance in organic chemistry.29-41 For a typical SN2 reaction, it is a backside attack of the substrate from the nucleophile with a Walden-inversion –

mechanism. On the PES of the title SN2 reaction F + CH3Cl constructed by Szabó –

and Czakó,42 there are two potential minima corresponding to the F ---CH3Cl and –

FCH3---Cl ion-dipole complexes in the entrance and exit channels respectively. In –

addition, for this PES, there is also a H-bonded minimum F ---HCH2Cl in the –

entrance channel which has a lower energy than the usual F ---CH3Cl pre-reactive complex. The relative energy barrier height with respect to the hydrogen-bonded complex is 4.8 kcal/mol, and the relative energy difference of the pre-active complex with respect to the hydrogen-bonded complex is 1.5 kcal/mol. The reaction energy barrier is below the asymptotic reactants by -12.8 kcal/mol with exoergic energy at -31.8 kcal/mol,42 which means this reaction has an early, negative energy barrier. Su et al. used a model potential to study the rate constants of the title reaction43 4

ACS Paragon Plus Environment

Page 5 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

assuming that the negative-barrier is not important to the reactivity. However, statistical theory study on rate constants by Hase’s group44 showed that negative-barrier did affect the rate constants. They also found in another classical trajectories study45 that the excitation of C-Cl stretch mode of CH3Cl had little effect on the reaction rates, but at high relative translational energies, the C-Cl excitation leads to a significant increase in the rate constants. Szabó and Czakó performed a quasi-classical trajectory(QCT) calculation on their full-dimensional potential energy surface46. They found that at low collision energies, the pre-reactive minimum –



F ---CH3Cl and H-bonded F ---HCH2Cl minimum in the entrance channel play a major role in the dynamics. At high collision energy, the direct rebound mechanism plays a dominate role for the dynamics47,48. They also found that the integral cross section decreases rapidly with increasing of the translational energy. Later, they updated their PES42 by including more reaction channels other than the backside attack and redid the quasi-classical calculation using the new PES. They found that the C-H symmetric stretching vibrational mode has only slight effect on the SN2 reactivity. Recently, a six degree-of-freedom (DOF) quantum dynamics and QCT calculation by Wang et al49 was carried to study the mode-specific reactivity on this reaction. The reaction probability ratios between the vibrational excited states and the ground state, as well as the integral cross section(ICS) of the ground state were calculated with both the quantum dynamics and QCT methods; however, the integral cross sections of the vibrational excited states were not calculated. In the present paper, we want to obtain not only the ICS of the ground state but also the ICSs of the 5

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 29

vibrational excited states; therefore we can investigate, under the equal amount of total energy, the energy efficacy on the reactivity; we want to find out whether the energy efficiency on reactivity also depends on the potential energy minima, besides the saddle point, on the PES for this reaction. If so, what roles do they play?

II. THEORY In this work, we performed a quantum dynamics calculation to study the reaction –



on the new updated F + CH3Cl PES by Szabó and Czakó.42 The F + CH3Cl → –

FCH3 + Cl SN2 reaction is a synchronized one bond broken and one bond formed mechanism, and the reaction also involves a Walden-inversion motion where the three H atoms move in sync and invert their configuration at the transition state. So in our quantum dynamics study, we included the three signature motions of this SN2 reaction: the bond-broken coordinate of the leaving group Cl to CH3, the bond-forming –

coordinate of the nucleophile F to the substrate CH3Cl, the three Hs move-in-phase coordinate, the umbrella motion, C-H3. Furthermore, we included the attacking angle –

between F and the CH3Cl to allow the nucleophile to attack the substrate from all directions. Since the QCT study by Szabó and Czakó42 shows that C-H symmetric stretching motion only severs as a spectator and has slight effect on the SN2 reactivity, thus it is not included in this model. Thus in this study, as seen in Figure 1, we use the reactant Jacobi coordinates to describe the system.

R is the distance from the nucleophile F– to the center of mass

CH3Cl to represent the attacking motion; r is the distance from the center of mass 6

ACS Paragon Plus Environment

Page 7 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

of CH3 to Cl to represent the leaving motion;

ρ is the distance from C atom to the

center of mass of the three hydrogen atoms to represent the Walden-inversion motion, thus here, the three hydrogen atoms are treated as one pseudo-atom X and move in phase;

γ is the attacking angle formed by the leaving coordinate r and the

attacking coordinate

R . So there are four degrees of freedom (4DOF) to describe the

title reaction's SN2 mechanism. The 4DOF Hamiltonian in the reactant Jacobi coordinates is written as,

(J − j h2 ∂ 2 H = − + 2 2µ ∂R 2µ R2 ∧

)

2

+ hr ( r

) + hρ ( ρ ) +

j

2

(1)

2µ r r r2

+ V4 D ( R, r , ρ , γ )

Here, the first term is the translational energy of the reaction system. Where

µ is the

reduced mass of the whole system and µ r is the reduced mass of CH3 and Cl; J

is

the total angular momentum operator of the system, and j is the rotational angular momentum operator of Cl-CH3. Jacobi angle θ and the torsion angle ϕ are fixed at their transition-state geometry. The vibrational reference Hamiltonians hr(r ) and hρ( ρ ) are given as,

hr ( r

)

= −

hρ ( ρ ) = −

h2 ∂ 2 + V1D ( r 2µr ∂r 2

)

h2 ∂ 2 + V1D ( ρ ) 2µρ ∂ρ 2

(2)

(3)

Where V1D(r ) and V1D( ρ ) are the one-dimensional effective potentials for r and

ρ coordinates by putting the reaction system in the reactant asymptotic region with other coordinates fixed at the equilibrium geometry. The split-operator method50 was employed here to propagate the wave packet for 7

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 29

the quantum scattering process. The time-dependent wavefunction was expanded in terms of the body-fixed rovibrational eigenfunctions in the reactant Jacobi coordinates. After the time-dependent wave function was propagated passing the transition state, the initial-state-selected reaction probability was extracted using the time-dependent reactive flux method.51,52 The initial-state-selected reaction probabilities of different

E) were added up to obtain initial-state-selected integral cross partial waves Pv0j 0( k0 J

section (ICS) σv 0j0(E ),

συ

0 j0

(E ) =

In the above Eq. (4),

1 π 2j0 + 1 k 2

∑ (2J

+ 1) PυJ j K

J

0 0

0

(E )

(4)

E is the collision energy, K 0 is the projection of J onto the

body-fixed z axis of the system, j 0 denotes the initial rotational quantum number,

ν 0 denotes the initial vibrational quantum number and k is the wavenumber,

k = ( 2uE)1/ 2 . The standard centrifugal sudden (CS) approximation53,54 was employed in calculation for J > 0 , since the computational cost of the title reaction is extremely expensive, the exact close-coupling(CC) calculation was not feasible. Nonetheless, our calculated ICS of the ground state has an excellent agreement with the ones of the 6DOF and QCT results(see Figure 2), which means that the CS calculation is a good approximation for the current system. For the translational coordinate R in the range from 3.3 to 25.0 bohr, we use 310 sine basis functions to expand the wave function, among these 110 are chosen to expand the wave function in the interaction region; we use 97 potential-optimized vibrational DVR points55 for the r coordinate in the range of 2.3 to 8.0 bohr, 12 potential-optimized vibrational DVR points for

ρ coordinate covering the range of 8

ACS Paragon Plus Environment

Page 9 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

0.0 to 1.5 bohr, and 110 spherical harmonic rotational functions for

γ . The

time-dependent wave packet is propagated with a time step of 15 atomic unit time (~ 0.36 fs) for a total propagation time of about 600,000 a.u. time ( ~ 14.5 ps). Note due –

to the long-range ion and dipole interaction between the nucleophile F and the substrate CH3Cl, a very long attacking coordinate up to 25.0 bohr was needed to describe the correct attacking motion. Furthermore, due to the two energy minima in the entrance channel, a very large 600,000 a.u. propagation time was needed to propagate the wave packet to obtain each converged reaction probability. On top of the above, reaction resonances were found in reaction probabilities with partial waves calculations, the J-shifting method56 cannot be used to obtain reliable ICSs. Therefore, with each integral cross section calculation, the reaction probability of every different partial wave up to J = 300 was needed to be calculated one by one. The computational cost was huge for this calculation.

III. RESULTS AND DISCUSSION The vibrational levels (ν r , ν ρ ) of ClCH3 assigned to their corresponding vibrational excitation states are listed in Table 1. Our calculated vibrational frequencies of Cl-CH3(1, 0) and C-H3(0, 1) are 740 cm-1 and 1486 cm-1. Their corresponding normal mode frequencies57 are 754 cm-1 and 1394 cm-1 on the new PES by Szabó and Czakó. In Figure 2, we compared our quantum integral cross section (ICS) of the ground state with the QCT result and the 6DOF result. This figure shows this reaction does 9

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 29

not have a reaction threshold due to a negative reaction barrier. The ICS decreases dramatically as the translational energy increases at relatively low translational energies. For example, the ICS of the ground state decreases about 89% from 1067.7 bohr2 to 110.9 bohr2 as the translational energy increases from 0.46 to 4.8 kcal/mol, then slowly decreases to 1.8 bohr2 at 15 kcal/mol. This means that the reactivity of the title reaction is dominated by the translational energies lower than 4.8 kcal/mol. Our quantum ICS of the ground state has an excellent agreement with the 6DOF quantum result and agrees very well with the QCT one. This comparison shows that our 4DOF quantum approach produces almost the same ICS as the 6DOF result does. Figure 3 shows the comparison of the ICSs among the ground state (0, 0), the first excited Cl-CH3 vibrational state (1, 0) and the first excited C-H3 vibrational state (0, 1). We calculated every reaction probability for different partial waves up to J = 300 to converge the three ICSs for translational energy up to 15 kcal/mol. All three ICSs decrease significantly as the translational energy increases until 4.8 kcal/mol. This indicates that the reactivity of the title reaction is indeed dominated by the translational energy lower than 4.8 kcal/mol. This translational energy 4.8 kcal/mol is the dividing point for the reactivity. In order to investigate the energy efficiency on reactivity, the ICS ratios between the vibrational excitation state (1, 0) and the ground vibrational state (0, 0), σ(1, 0)/σ(0, 0), and between (0, 1) and (0, 0), σ(0, 1)/σ(0, 0), are shown in Figure 4A and 4B on the basis of an equal amount of total energy. Figure 4 demonstrates that both ratios decrease to reach a minimum at 4.8 kcal/mol and increase at relative high 10

ACS Paragon Plus Environment

Page 11 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

energies after 4.8 kcal/mol. This translational energy 4.8 kcal/mol serves as another dividing point again, here as the dividing point on two different energy efficiency roles on reactivity. Namely, the energy efficiency on reactivity has been distinguished on two sides of this dividing point at 4.8 kcal/mol. This dividing energy point has to be associated with some special characters on the PES. We found that the dividing energy point, 4.8 kcal/mol, is exactly the relative energy barrier height of the negative –

barrier (-12.8 kcal/mol) with respect to the lower energy minimum, the F ---HCH2Cl hydrogen-complex(-17.6 kcal/mol), in the entrance channel.42 This means the energy efficacy at the left side of 4.8 kcal/mol is associted with translational energies smaller –

than the relative barrier height with respect to the F ---HCH2Cl hydrogen-bond complex, and the right side of 4.8 kcal/mol is associted with translational energies –

larger than the relative barrier height with respect to the F --- HCH2Cl hydrogen-bond complex. Thus the energy efficacy on reactivity for tranlational energies lower than 4.8 kcal/mol (in Figure 4), or the dominant reactivity for tranlational energies lower than 4.8 kcal/mol (in Figure 2), depends on the relative barrier height between the transition state and the hydrogen-bond minimum, which means the dominant reactivity here is related to this minimum and the transition state. That is, the biggest reactivity takes palces when the reactants go through the wells in the entrance channel first then scale the transtition state. The dominance of low collision energies by forming the pre-reaction complex has also been found in the direct dynamics classical –

trajectory simulation study on the similar Cl + CH3I reaction by Zhang et al.58 In Szabó and Czakó's QCT study,46 they also found the trajectories indeed can trap in the 11

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 29

two minima in the entrance channel to form long-lived complexes. The indirect reaction mechanism that reactants trapped in the pre-reaction complexes has also been observed in the studies of F− + CH3I40 and OH− + CH3I41 since the PESs for F− + CH3I and OH− + CH3I are quite similar to that for F− + CH3Cl. The above studies show that although these reactions have negative barriers, the reactants need to go through the potential energy minima to for the pre-reaction complexes to let reactions occur, and low-collision-energy reactivities dominate these reactions. Note at the translational energy 1.5 kcal/mol, the values of the two ICS ratios are equilvalent at 3.0. For translational energies lower than 1.5 kcal/mol, all the σ(0, 1)/σ(0, 0) values in Figure 4B are bigger than the σ(1, 0)/σ(0, 0) values in Figure 4A. For translational energies larger than 1.5 kcal/mol, the situation is reversed as all the σ(1, 0)/σ(0, 0) values in Figure 4A are bigger than the σ(0, 1)/σ(0, 0) values in Figure 4B. This energy value 1.5 kcal/mol at this equilvalent ratio point happenes to be the –

energy difference of the pre-reactive complex F ---CH3Cl (-16.1 kcal/mol) with –

respect to the hyrogen-bond complex F ---HCH2Cl(-17.6kcal/mol) in the entrance channel.42 Based on the above observed features, there are two dividing points 4.8 kcal/mol and 1.5 kcal/mol at the relative low translational energies: one is the relative barrier height, 4.8 kcal/mol, of the transition state with respect to the hyrogen-bond –

complex F ---HCH2Cl, the other is the relative energy height, 1.5 kcal/mol, of the –

pre-reactive complex F ---CH3Cl with respect to the hyrogen-bond complex –

F ---HCH2Cl. Therefore the energy effacicy on reactivity, for the translational energies lower than 4.8 kcal/mol (as seen in Figure 2), is relative to the PES features 12

ACS Paragon Plus Environment

Page 13 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

of the transition state and the pre-reacitve complex with respect to the lowest hydrogen-bonded complex mininum respectively. In other words, the donimant reactivity(in Figure 2), for the translational energies lower than 4.8 kcal/mol, is controlled by the two energy minina and the transition state together. As mention above, the dominance of the indirect reaction mechanisms at low collision energies by –

forming the pre-reaction complex has also been found in the similar reaction Cl + CH3I by Zhang et al.58 Furthermore, for the translational energies lower than 1.5 kcal/mol, the C-H3 vibration is more efficient than the CH3-Cl vibration in promoting this reaction. This –

indicates the C-H3 vibration motion favors the formation of the F --HCH2Cl –

hydrogen-bond complex which is 1.5 kcal/mol lower than the F ---CH3Cl complex. For the translational energies larger than 1.5 kcal/mol, the CH3-Cl vibration is more effecitent than the C-H3 vibration. It is understandable because the CH3-Cl vibration –

favors the reactants to form F ---CH3Cl complex whose energy is 1.5 kcal/mol higher –

than the hydrogen complex F ---HCH2Cl. When the collision energy is lower than 1.5 –

kcal/mol to form the F ---HCH2Cl hydrogen complex, this is the indirect mechanism –

as F has to hydrogen-bond with one H first then rotate itself to face and attack the backside of the C atom. For energies larger than 1.5 kcal/mol up to 4.8 kcal/mol, the CH3-Cl vibration is more efficient than the C-H3 vibration since it mainly forms the –

F ---CH3Cl pre-reactive complex. This is the indirect rebound reaction mechanism –

since the nucleophile F needs to form the pre-reactive complex with the substrate CH3Cl, then attack C center atom directly, or it is called the backside attack 13

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 29

mechanism described in textbooks for the SN2 reactions. Overall, on both sides of the minimum, the vibrational motions of Cl-CH3 and C-H3 are more efficient than the transiational motion in enhancing the reactivity. As for the translational energies below the relative barrier height, Figure 4 shows that both vibrational stretching of CH3-Cl and C-H3 are more important than the translational energy in enhancing the reactivity; however, as the translational energy increases, the importance of these two vibrations in enhancing reactivity decreases; Furthermore, Figure 2 shows that the ICSs decrease dramatically before 4.8 kcal/mol; these two features are due to the fact that, as translational energy increases, the reactants will not have enough time to orientate themselves to form the two minima complexes in the entrance channel. The energy efficacy comparison provides direct evidence here that the two minima in the entrance channel play an essential role on reactivity at the lower translational energies: they determine the energy efficacy on reactivity at the relative low translational energies. For the translational energies larger than the relative reaction barrier, 4.8 –

kcal/mol, the reactants can skip the reactant channel minima and the nucelophile F

can attack the C atom from the backside directly to let the reaction occure, thus it's a direct rebound mechanism. In addition, since the translational energy at this part is larger than the relative barrier height, so the reaction can occur without forming the –

transition-state complex, and the nucelophile F can approch the side of the substrate to directly strip CH3 away. Hence there are two reaction mechanisms occuring simultaneously at this energy range: the direct rebound and strip away mechanisms. 14

ACS Paragon Plus Environment

Page 15 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

For energy higher than 4.8 kcal/mol, both the CH3-Cl and C-H3 stretching motions become more efficient than at relative lower translational energy because these two –

vibrations make the substrate a larger target to be approached, and thus make the F

directly attack C atom or strip away CH3 easier. This is also the reason, in the QCT calculations by Szabó and Czakó,42,46 they found the stripping mechasiman happens for large impact parameters. Xie et al. have also discussed both the direct and indirect – – – – – mechanisms in the similar reactions of X (X = F , Cl , OH ) + CH3I.47

Note both the Cl-CH3 vibration and the C-H3 vibration are more efficient than the translational motion in enhancing the reactivity at relative high translational energies; however, the Cl-CH3 ratio is larger than the one of C-H3, which means the Cl-CH3 vibration plays a more effective role than the C-H3 vibration on promoting this reaction. The Cl-C stretch excitation promoting reaction for a high translational energy has also been found in the trajectory study45. That the Cl-CH3 vibration is more effective than the C-H3 vibration on promoting this reaction has been also observed in the 6DOF quantum dynamics calculation;49 however, the 6DOF ICSs of these two vibrtional excited states were not calculated in that study, the conclusion was drawn from the reaction probability comparison. There are also qunatum dynamics studies that have been done on the similiar, –



non-symmetric SN2 reactions Cl + CH3Br59,60 and Cl + CH3I61. Both the 4DOF –

time-independent quantum studies of Cl + CH3Br and the 3DOF time-dependent –

quantum dynamics study of the Cl + CH3I indicate that, in general, the C-H3 umbrella mode and the C-X(Cl, Br, I) stretching mode in the substrate, similar as in 15

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 29

the title reaction, can both enhance the reactivities of these non-symmetric SN2 reactions.

IV. CONCLUSIONS –



Overall, for the F + CH3Cl → FCH3 + Cl SN2 reaction, it is not the absolute negative barrier, but the relative barrier and the minima in the entrance channel that determine the energy efficiency on the reactivity. The Cl-CH3 vibrational motion is more efficient than the C-H3 vibrational motion in promoting this reaction. For the Cl-CH3 vibrational motion, the ratio of the ICS is larger than one for the whole energy range we studied here, which means the Cl-CH3 vibrational motion is more efficient than the translational motion in enhancing the reactivity, although this reaction has an early barrier. Around the relative barrier height ~ 4.8 kcal/mol, the translational energy almost has the same efficiency as vibrational motions. Nevertheless, the ICSs are dominated by the translational energies lower than 4.8 kcal/mol, which means that the reactivity is most effective when involving the two potential minima in the entrance channel. Or in other words, the two minima on the PES have an dominate effect on the reactivity. Therefore, the energy efficacy is determined by not only the relative transition state, but also more importantly the two minima in the entrance channel for this reaction. Recent studies show that the Polanyi rules cannot be simply extended to poly-atomic reactions for more than three-atom systems. In these studies, the energy efficacy was considered solely based on the locations of the transition states. However, our results on the energy efficacy based on the equal amount of total 16

ACS Paragon Plus Environment

Page 17 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

energy show that potential energy minima in the entrance channel also play a significant role on the reactivity. Thus, for the poly-atomic reactions with minima on –



their PESs, at least for the SN2 reactions as X + CH3Y → XCH3 + Y , one needs to consider the potential energy minimum (minima) to get a general energy efficacy rule on reactivity. Nonetheless, we think more theoretical studies, as well as experimental studies, are needed to reveal the potential minima effects on the energy efficacy in promoting reactivity for poly-atomic reaction systems.

ACKNOWLEDGEMENTS We thank Prof. Czakó for sending us the PES of the title reaction. This work is supported by the National Natural Science Foundation of China (Grant No. 11374194) and Taishan Scholarship Fund from Shandong province. The computation was carried out at the Shenzhen Supercomputer Center in China.

17

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 29

References (1) Polanyi, J. C. Some concepts in reaction dynamics. Science 1987, 236, 680-690. (2) Sanson, J.; Corchado, J. C.; Rangel, C.; Espinosa-Garcia, J. Quasiclassical trajectory calculations comparing the reactivity and dynamics of symmetric and asymmetric stretch and the role of the bending mode excitations of methane in the Cl + CH4 reaction. J. Chem. Phys. 2006, 124, 074312. (3) Meng, F. B.; Yan, W.; Wang, D.Y. Quantum dynamics study of the Cl + CH4 → HCl + CH3 reaction: reactive resonance, vibrational excitation reactivity, and rate constants. Phys. Chem. Chem. Phys. 2012, 14, 13656-13622. (4) Zhang, Z.; Zhou, Y.; Zhang, D. H.; Czakó, G.; Bowman, J. M. Theoretical study of the validity of the Polanyi rules for the late-barrier Cl + CHD3 reaction. J. Phys. Chem. Lett. 2012, 3, 3416-3419. (5) Wang, F.; Lin, J. S.; Cheng, Y.; Liu, K. Vibrational enhancement factor of the Cl + CHD3 (v1=1) reaction: rotational-probe effects. J. Phys. Chem. Lett. 2013, 4, 323-327. (6) Czakó, G.; Bowman, J. M. Accurate ab initio potential energy surface, thermochemistry, and dynamics of the Cl (2P, 2P3/2) + CH4→ HCl + CH3 and H + CH3Cl reactions. J. Chem. Phys. 2012, 136, 044307. (7) Li, J.; Jiang, B.; Guo, H. Reactant vibrational excitations are more effective than translational energy in promoting an early-barrier reaction F + H2O → HF + OH. J. Am. Chem. Soc. 2013, 135, 982-985. (8) Jiang, B.; Guo, H. Control of mode/bond selectivity and product energy disposal by the transition state: X + H2O (X= H, F, O (3P), and Cl) reactions. J. Am. Chem. Soc. 2013, 135, 15251-15256. (9) Wang, D.; Czakó, G. Quantum dynamics study of the F + CH4 → HF + CH3 reaction on an ab initio potential energy surface. J. Chem. Phys. A 2013, 117, 7124-7130. (10) Zhang, W.; Kawamata, H.; Liu. K. CH stretching excitation in the early barrier F + CHD3 reaction inhibits CH bond cleavage. Science 2009, 325, 303-306. 18

ACS Paragon Plus Environment

Page 19 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(11) Czakó, G.; Bowman, J. M. CH stretching excitation steers the F atom to the CD bond in the F + CHD3 reaction. J. Am. Chem. Soc. 2009, 131, 17534-17535. (12) Czakó, G.; Bowman, J. M. Quasiclassical trajectory calculations of correlated product distributions for the F + CHD3 (v1=0,1) reactions using an ab initio potential energy surface. J. Chem. Phys. 2009, 131, 244302. (13) Pan,

H.;

Liu,

K.

Communication:

Imaging

the

effects

of

the

antisymmetric-stretching excitation in the O (3P) + CH4 (v3= 1) reaction. J. Chem. Phys.2014, 140, 191101. (14) Zhang, B.; Liu, K. How active is the bend excitation of methane in the reaction with O (3P)? J. Phys. Chem. A 2005, 109, 6791-6795. (15) Liu, R.; Yang, M.; Czakó, G.; Bowman, J. M.; Li, J.; Guo, H. Mode selectivity for a “central” barrier reaction: Eight-dimensional quantum studies of the O (3P) + CH4 → OH + CH3 reaction on an ab initio potential energy surface. J. Phys. Chem. Letts. 2012, 3, 3776-3780. (16) Yan, W.; Meng, F.; Wang, D. Quantum dynamics study of vibrational excitation effects and energy requirement on reactivity for the O + CD4/CHD3 → OD/OH + CD3 reactions. J. Phys. Chem. A 2013, 117, 12236-12242. (17) Wang, F.; Liu, K. Enlarging the reactive cone of acceptance by exciting the C–H bond in the O (3P) + CHD3 reaction. Chem. Sci. 2010, 1, 126-133. (18) Zhang, Z.; Chen, J.; Liu, S.; Zhang, D. H. Accuracy of the centrifugal sudden approximation in the H + CHD3 → H2 + CD3 reaction. J. Chem. Phys. 2014, 140, 224304. (19) Welsch, R.; Manthe, U. Communication: Ro-vibrational control of chemical reactivity in H + CH4 → H2 + CH3: Full-dimensional quantum dynamics calculations and a sudden model. J. Chem. Phys. 2014, 141, 051102. (20) Zhou, Y.; Wang, C.; Zhang, D. H. Effects of reagent vibrational excitation on the dynamics of the H + CHD3 → H2 + CD3 reaction: A seven-dimensional time-dependent wave packet study. J. Chem. Phys. 2011, 135, 024313. (21) Welsch, R.; Manthe, U. The role of the transition state in polyatomic reactions: Initial state-selected reaction probabilities of the H + CH4 → H2 + CH3 reaction. J. 19

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 29

Chem. Phys. 2014, 141, 174313. (22) Jiang, B.; Guo, H. Mode specificity, bond selectivity, and product energy disposal in X + CH4/CHD3 (X= H, F, O (3P), Cl, and OH) hydrogen abstraction reactions: perspective from sudden vector projection model. J. Chin. Chem. Soc. 2014, 61, 847-859. (23) Li, J.; Dawes, R.; Guo, H. An ab initio based full-dimensional global potential energy surface for FH2O (X2A′) and dynamics for the F + H2O → HF + HO reaction. J. Chem. Phys. 2012, 137, 094304. (24) Czakó, G.; Bowman, J. M. Dynamics of the O (3P) + CHD3 (vCH= 0, 1) reactions on an accurate ab initio potential energy surface. Proc. Natl. Acad. Sci.U.S.A. 2012, 109, 7997-8001. (25) Czakó, G.; Bowman, J. M. Dynamics of the reaction of methane with chlorine atom on an accurate potential energy surface. Science 2011, 334, 343-346. (26) Liu, J.; Song, K.; Hase, W. L.; Anderson, S. L. Direct dynamics trajectory study of vibrational effects: Can polanyi rules be generalized to a polyatomic system? J. Am. Chem. Soc. 2004, 126, 8602-8603. (27) Jiang, B.; Guo, H. Mode specificity, bond selectivity, and product energy disposal in X + CH4/CHD3 (X= H, F, O (3P), Cl, and OH) hydrogen abstraction reactions: perspective from sudden vector projection model. J. Chin. Chem. Soc. 2014, 61, 847-859. (28) Guo, H.; Jiang, B. The sudden vector projection model for reactivity: mode specificity and bond selectivity made simple. Acc. Chem. Res. 2014, 47, 3679-3685. (29) Olmstead, W. N.; Brauman, J. I. Gas-phase nucleophilic displacement reactions. J. Am. Chem. Soc. 1977, 99, 4219-4228. (30) Hase, W. L. Simulations of gas-phase chemical reactions: Applications to SN2 nucleophilic substitution. Science 1994, 11, 999-1002. (31) Chabinyc, M. L.; Craig, S. L.; Regan, C. K.; Brauman, J. I. Gas-phase ionic reactions: Dynamics and mechanism of nucleophilic displacements. Science 1998, 279, 1882-1886. 20

ACS Paragon Plus Environment

Page 21 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(32) Peslherbe, G. H.; Wang, H.; Hase, W. L. Unimolecular dynamics of Cl−...CH3Cl intermolecular complexes formed by Cl− + CH3Cl association. J. Chem. Phys. 1995, 102, 5626-5635. (33) Wang, H.; Hase, W. L. Statistical rate theory calculations of the Cl− + CH3Br → ClCH3 + Br− rate constant versus temperature, translational energy, and H (D) isotopic substitution. J. Am. Chem. Soc. 1995, 117, 9347-9356. (34) Mikosch, J.; Otto, R.; Trippel, S.; Eichhorn, C.; Weidemuller, M.; Wester, R. Inverse temperature dependent lifetimes of transient SN2 ion-dipole complexes. J. Phys. Chem. A 2008, 112, 10448-10452. (35) Mikosch, J.; Trippel, S.; Eichhorn, C.; Otto, R.; Lourderaj, U.; Zhang, J. X.; Wester, R. Imaging nucleophilic substitution dynamics. Science 2008, 319, 183-186. (36) Wester, R.; Bragg, A. E.; Davis, A. V.; Neumark, D. M. Time-resolved study of the symmetric SN2-reaction I− + CH3I. J. Chem. Phys. 2003, 119, 10032-10039. (37) Vande Linde, S. R.; Hase, W. L. Trajectory studies of SN2 nucleophilic substitution. I. Dynamics of Cl− + CH3Cl reactive collisions. J. Chem. Phys. 1990, 93, 7962-7980. (38) Vande Linde, S. R.; Hase, W. L. A direct mechanism for SN2 nucleophilic substitution enhanced by mode-selective vibrational excitation. J. Am. Chem. Soc. 1989, 111, 2349-2351. (39) Manikandan,P.; Zhang, J.; Hase, W. L. Chemical dynamics simulations of X– + CH3Y → XCH3 + Y– Gas-phase SN2 nucleophilic substitution reactions. Nonstatistical dynamics and nontraditional reaction mechanisms. J. Phys. Chem. A 2012, 116, 3061-3080. (40) Mikosch, J.; Zhang, J.; Trippel, S.; Eichhorn, C.; Otto, R.; Sun, R.; Wester, R. Indirect dynamics in a highly exoergic substitution reaction. J. Am. Chem. Soc. 2013, 135, 4250-4259. (41) Xie, J.; Sun, R.; Siebert, M. R.; Otto, R.; Wester, R.; Hase, W. L. Direct dynamics simulations of the product channels and atomistic mechanisms for the OH– + CH3I reaction. Comparison with experiment. J. Phys. Chem. A 2013, 117, 7162-7178. 21

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 29

(42) Szabó, I.; Czakó, G. Revealing a double-inversion mechanism for the F− + CH3Cl SN2 reaction. Nat. Commun. 2015, 6, 5972. (43) Su, T.; Morris, R. A.; Viggiano, A. A.; Paulson, J. F. Kinetic energy and temperature dependences for the reactions of fluoride with halogenated methanes: experiment and theory. J. Phys. Chem. 1990, 94, 8426-8430. (44) Wang, H.; Hase, W. L. Kinetics of F− + CH3Cl SN2 nucleophilic substitution. J. Am. Chem. Soc. 1997, 119, 3093-3102. (45) Su, T.; Wang, H.; Hase, W. L. Trajectory studies of SN2 nucleophilic substitution. 7. F− + CH3Cl → FCH3 + Cl−. J. Phys. Chem. A 1998, 102, 9819-9828. (46) Szabó, I.; Császár, A. G.; Czakó, G. Dynamics of the F− + CH3Cl → Cl− + CH3F SN2 reaction on a chemically accurate potential energy surface. Chem. Sci. 2013, 4, 4362-4370. (47) Xie, J.; Otto, R.; Mikosch, J.; Zhang, J.; Wester, R.; Hase, W. L. Identification of atomic-level mechanisms for gas-phase X– + CH3Y SN2 reactions by combined experiments and simulations. Acc. Chem. Res. 2014, 47, 2960-2969. (48) Zhang, J.; Mikosch, J.; Trippel, S.; Otto, R.; Weidemüller, M.; Wester, R.; Hase, W. L. F− + CH3I→ FCH3 + I− reaction dynamics. Nontraditional atomistic mechanisms and formation of a hydrogen-bonded complex. J. Phys. Chem. Lett. 2010, 1, 2747-2752. (49) Wang, Y.; Song, H.; Szabó, I.; Czakó, G.; Guo, H.; Yang, M. Mode-specific SN2 reaction dynamics. J. Phys. Chem. Lett., 2016, 7, 3322-3327. (50) Fleck, J. A.; Morris, J. R.; Feit, M. D. Time-dependent propagation of high energy laser beams through the atmosphere. Appl. Phys., 1976, 10, 129-160. (51) Zhang, D. H.; Zhang, J. Z. H. Full‐dimensional time‐dependent treatment for diatom–diatom reactions: The H2 + OH reaction. J. Chem. Phys. 1994, 101, 1146-1156. (52) Zhang, D. H.; Zhang, J. Z. H. Quantum reactive scattering with a deep well: Time ‐dependent calculation for H + O2 reaction and bound state characterization for HO2. J. Chem. Phys. 1994, 101, 3671-3678. (53) McGuire, P.; Kouri, D. J. Quantum mechanical close coupling approach to 22

ACS Paragon Plus Environment

Page 23 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

molecular collisions. jz‐conserving coupled states approximation. J. Chem. Phys., 1974, 60, 2488-2499. (54) Pack, R. T. Space‐fixed vs body‐fixed axes in atom‐diatomic molecule scattering. Sudden approximations. J. Chem. Phys. 1974, 60, 633-639. (55) Echave, J.; Clary, D. C. Potential optimized discrete variable representation. Chem. Phys. Lett. 1992, 190, 225-230. (56) Bowman, J. M. Reduced dimensionality theory of quantum reactive scattering. J. Phys. Chem., 1991, 95, 4960-4968. (57) Szabó, I.; Czakó, G. Revealing a double-inversion mechanism for the F− + CH3Cl SN2 reaction. Nat. Commun. 2015, 6, 5972. (Data in Supporting Information) (58) Zhang, J.; Lourderaj, U.; Sun, R.; Mikosch, J.; Wester, R.; Hase, W. L. Simulation studies of the Cl− + CH3I SN2 nucleophilic substitution reaction: Comparison with ion imaging experiments. J. Chem. Phys. 2013, 138, 114309. (59) Schmatz, S. Four-mode quantum calculations of resonance states in complex-forming bimolecular reactions: Cl− + CH3Br. J. Chem. Phys. 2005, 122, 234306. (60) Hennig, C.; Schmatz, S. Four-dimensional quantum study on exothermic complex-forming reactions: Cl− + CH3Br → ClCH3 + Br−. J. Chem. Phys. 2005, 122, 234307. (61) Kowalewski, M.; Mikosch, J.; Wester, R.; de Vivie-Riedle, R. Nucleophilic substitution dynamics: Comparing wave packet calculations with experiment. J. Phys. Chem. A 2014, 118, 4661-4669.

23

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 29

Table 1. Initial vibrational states (ν r ,ν ρ) of ClCH3. The energy levels Ea and E – are relative to the energy of the F + CH3Cl asymptote and the ground vibrational

state of CH3Cl, respectively.

State No.

(ν r , ν ρ )

1

(0, 0)

1119

0

2

(1, 0)

1859

740

3

(2, 0)

2586

1467

4

(0, 1)

2605

1486

5

(3, 0)

3305

2186

6

(1, 1)

3339

2220

7

(4, 0)

4015

2896

8

(2, 1)

4060

2941

9

(0, 2)

4091

2972

Ea(cm-1)

E(cm-1)

24

ACS Paragon Plus Environment

Page 25 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 1. Reactant Jacobi coordinates for the





F + CH3Cl → FCH3 + Cl SN2

reaction.

25

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 29

Figure 2. The comparison of the integral cross sections of the ground state among our 4DOF quantum result, the QCT result[Ref. 49], and 6DOF quantum result[Ref. 49].

26

ACS Paragon Plus Environment

Page 27 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 3. The integral cross sections of the (0, 0), (1, 0), and (0, 1) states as a function of (A) translational energy and (B) total energy. In order to show the ICSs' difference at relative high translational energies, a logarithmic plot of the ICSs is also inserted in Figure A. 27

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 29

Figure 4. The ICS ratios for σ(1, 0)/σ(0,0) and σ(0, 1)/σ(0, 0) against the translational energy. The vibrational energy difference ∆E1 between the first Cl-CH3 excited state(1, 0) and the ground state(0, 0) is 2.1 kcal/mol. The vibrational energy difference

∆E 2 between the first C-H3 exited state (0, 1) and the ground state (0, 0) is 4.2 kcal/mol. 28

ACS Paragon Plus Environment

Page 29 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

TOC graphic

29

ACS Paragon Plus Environment