Skip to main content
Philosophical transactions. Series A, Mathematical, physical, and engineering sciences logoLink to Philosophical transactions. Series A, Mathematical, physical, and engineering sciences
. 2017 Mar 20;375(2092):20160204. doi: 10.1098/rsta.2016.0204

Perspective: chemical dynamics simulations of non-statistical reaction dynamics

Xinyou Ma 1, William L Hase 1,
PMCID: PMC5360902  PMID: 28320906

Abstract

Non-statistical chemical dynamics are exemplified by disagreements with the transition state (TS), RRKM and phase space theories of chemical kinetics and dynamics. The intrinsic reaction coordinate (IRC) is often used for the former two theories, and non-statistical dynamics arising from non-IRC dynamics are often important. In this perspective, non-statistical dynamics are discussed for chemical reactions, with results primarily obtained from chemical dynamics simulations and to a lesser extent from experiment. The non-statistical dynamical properties discussed are: post-TS dynamics, including potential energy surface bifurcations, product energy partitioning in unimolecular dissociation and avoiding exit-channel potential energy minima; non-RRKM unimolecular decomposition; non-IRC dynamics; direct mechanisms for bimolecular reactions with pre- and/or post-reaction potential energy minima; non-TS theory barrier recrossings; and roaming dynamics.

This article is part of the themed issue ‘Theoretical and computational studies of non-equilibrium and non-statistical dynamics in the gas phase, in the condensed phase and at interfaces’.

Keywords: chemical reaction dynamics, non-statistical dynamics, post-transition state dynamics, non-intrinsic reaction coordinate dynamics

1. Introduction

Statistical theories in chemical dynamics and kinetics are theoretical models whose assumptions lead to the use of statistical mechanics to calculate properties for chemical reactions, such as rate constants and product energy partitioning distributions [1]. Transition state theory (TST) is derived using classical mechanics and is based on two fundamental assumptions: (i) the reactants have Boltzmann energy distributions and (ii) once a classical trajectory crosses the transition state (TS) towards products, it does not turn around and recross the TS forming reactants (i.e. the no-recrossing assumption of TST). TST is a canonical classical statistical mechanical model for chemical kinetics. Attempts to derive a quantum TST have been unsuccessful [2,3], and Miller has stated ‘there is no rigorous quantum version of transition-state theory’ [2]. The widely used model quantum TST results by replacing the classical partition functions for the TS and reactant(s) with their quantum counterparts.

Rice–Ramsperger–Kassel–Marcus (RRKM) unimolecular rate theory [47] is based on the fundamentals of microcanonical statistical mechanics; in the mass spectrometry community, RRKM theory is often referred to as quasi-equilibrium theory [6]. The fundamental assumption of the theory is that a microcanonical ensemble of vibrational states is maintained for the unimolecular reactant as it decomposes to products. RRKM is a TST and the RRKM rate constant is derived using classical mechanics. It also has the no-recrossing of the TS assumption. What is known as quantum RRKM theory results by replacing the sum of states at the TS and the density of states of the unimolecular reactant with their quantum counterparts. RRKM theory is a microcanonical statistical mechanical model for unimolecular kinetics.

An important component of statistical models of chemical reactions is the assumed motion on the potential energy surface (PES) before products are formed [8]. The nature of the dynamics on the PES, before reaching a rate-controlling TS, does not affect the reaction kinetics, but the dynamics after passing this TS are critical. They are referred to as ‘post-TS’ dynamics [9,10]. To predict attributes of chemical reactions after passing this TS, non-TST models are needed. Post-TS dynamics are determined by features of the PES, and they include branching between different product channels and reaction pathways and partitioning of the available energy to the reaction products.

The PES in this post-transition state domain may have a very ‘rough’ landscape, with multiple potential energy minima, reaction pathways, low energy barriers, etc., connecting the rate-controlling TS to multiple product channels. The important question is ‘what are the post-TS dynamics?’ A statistical model is often used to model the post-TS dynamics. For this model, potential energy minima, barriers and different product channels are connected via reaction paths determined by following their intrinsic reaction coordinates [11]. For the gas phase, where the energy of the reactive system is conserved, the statistical model uses RRKM theory to determine the time the system spends in each potential energy minima and the probabilities for transitions between minima and forming products. The statistical model for gas-phase product energy partitioning is phase space theory (PST) [1215] or a modified PST [16,17].

An intrinsic reaction coordinate (IRC) is the path of steepest descent from a TS to a potential energy minimum [11]. This motion results from setting the velocity of each atom to zero after each infinitesimal step. As energy is constant for gas-phase events, there is not a gas-phase trajectory that has the IRC motion. The IRC provides well-defined structures between stationary points (i.e. TSs, potential energy minima, reactants and products) on the PES.

Non-statistical dynamics are properties of chemical reactions which lead to violations of the fundamental assumptions of statistical theories, and/or the assumed statistical dynamics for motion on the PES. The following are important illustrative examples of non-statistical reaction dynamics, with chemical dynamics simulations by the author's research group emphasized. Extensive non-statistical dynamics have been observed, both experimentally and computationally, for X + CH3Y → XCH3 + Y SN2 bimolecular substitution reactions [18] and a number of the examples discussed below are for these reactions. As illustrated in figure 1, the PES for these reactions have X‐‐‐CH3Y pre- and XCH3‐‐‐Y post-reaction complexes and a [X‐‐CH3‐‐Y] central barrier TS.

Figure 1.

Figure 1.

Potential energy curve for the Cl + CH3I → ClCH3 + I reaction. The energies are given by MP2/ECP/aug-cc-pVDZ theory. Adapted from [8,18].

2. Illustrative examples of non-statistical dynamics

(a). Post-transition state dynamics

If a rate-controlling TS provides an accurate rate constant for the chemical reaction (i.e. TS recrossing is unimportant), an ensemble of trajectories may be initiated at the TS and propagated to products to study the post-TS dynamics [10]. For constant temperature unimolecular or bimolecular reactions, the initial conditions for these trajectories are chosen in accord with quantum TST. For a constant energy unimolecular reaction, for which the unimolecular dynamics are statistical and in accord with RRKM theory, to study the post-TS dynamics initial conditions are chosen at the TS as specified by quantum RRKM theory [19]. Quasi-classical initial conditions are chosen to sample the TS's quantum mechanical energy levels and explicitly include the TS zero-point energy (ZPE). The following are examples of non-statistical post-TS dynamics.

(i). Bifurcations and branching between exit-channel pathways

For a reaction whose PES has a bifurcation after passing the rate-controlling TS, chemical dynamics simulations are critical for predicting the non-statistical product branching ratio [2025]. The IRC [11] leads to a valley-ridge inflection (VRI) point and then dynamics are needed to determine the branching between the two product channels. A model PES has been developed to study non-statistical dynamics associated with PESs with reaction path bifurcations and VRI points [24]. Simulations on this PES [24] indicate that, in the gas phase, dynamics in the vicinity of the VRI are insignificant in determining the product branching; however for a reactive system with a strongly interacting solvent, and extensive energy transfer from the reactive system, the VRI may play an important role. Potential energy release to the solvent, in moving from the TS to the VRI, will tend to constrain the motion near the vicinity of the IRC and, thus, make dynamics in the vicinity of the VRI more important.

(ii). Product energy partitioning

Repulsive potential energy release. If there is repulsive potential energy release as the reactive system moves directly from the rate-controlling TS to products, statistical theory cannot be used to accurately model the product energy partitioning. Product energy partitioning for such a case is illustrated by the direct dynamics simulations for C2H5F → HF + C2H4 dissociation [19,26,27]. The simulations were performed using MP2 theory with both the 6-31G* and 6-311++G** basis sets, and the results are compared with experiments of Setser, Wittig and co-workers [28,29]. The experimental 0 K potential energy release with ZPE included, Ero, in going from the TS to products is 49.0 ± 2.0 kcal mol−1, while the MP2 values with the 6-31G* and 6-311++G** basis sets are 45.3 and 50.6 kcal mol−1, respectively.

An important comparison with experiment is the partitioning of the available product energy to HF vibration. For the experiments, excess energies at the TS (Evt) of 32 and 42 kcal mol−1 were investigated. For the latter Evt, the experiments show that 15% of the available energy is partitioned to HF vibration. The MP2/6-311++G** simulation gives 14% and nearly the same result, while the MP2/6-31G* simulation gives 8%. A more detailed comparison with experiment is the population distribution, P(n), of the HF vibrational levels, which is shown in figure 2. The MP2/6-311++G** simulations are in very good agreement with both the Evt of 32 and 42 kcal mol−1 experiments. Their principal difference is the small P(1)/P(0) inversion found from the simulations.

Figure 2.

Figure 2.

Populations of the HF vibrational states for C2H5F → HF + C2H4 dissociation, with different amounts of vibration/reaction coordinate energy Evt in the C2H5F transition state. Open squares, results of the MP2/6-311++G** simulations; filled circles, results of the MP2/6-31G* simulations from [17]; filled triangles, experimental results. The experimental results for Evt of 32 and 42 kcal mol−1 are from [21] and [20], respectively. Adapted from [26].

Also of interest are the percentages for partitioning to the products relative translation, rotation and vibration energies, obtained from the simulations but not the experiments. The results are given in table 1. The majority of the product degrees of freedom are for C2H4 vibration and for a statistical model the majority of the product energy would be partitioned to C2H4 vibration. By contrast, as shown in table 1, the majority of the energy is distributed to HF + C2H4 relative translation, arising from non-statistical, repulsive potential energy release in moving from the TS to products. In comparing the results of the two basis sets, 6-311++G** partitions slightly less to relative translation and more to HF vibration, with the other energy partitionings similar for the two basis sets.

Table 1.

Average product energy partitioning for C2H5F → HF + C2H4 direct dynamicsa.

MP2/6-31G*
MP2/6-311++G**
product energy Evt= 3.45 27 42 3.45 32 42
rel trans 75.4 56.8 49.7 67.8 48.2 46.2
C2H4 vib 6.1 24.2 31.5 6.8 26.2 29.3
C2H4 rot 4.6 5.9 6.0 5.1 5.5 6.2
HF vib 10.5 8.7 7.9 16.9 15.8 14.0
HF rot 3.4 4.4 4.9 3.4 4.3 4.3

aThe average energy partitioning is given in per cent. Evt is in kcal mol−1.

Following a model proposed by Zamir & Levine [30], the product energy partitioning was analysed versus Evt to determine the percentages of the potential energy release, Ero, distributed to the different product energies. For the MP2/6-31G* simulations, the percentages partitioned to relative translation, C2H4 vibration, C2H4 rotation, HF vibration and HF rotation are 81, 0, 5, 11 and 3%, respectively. For the MP2/6-311++G** simulations, these percentages are 73, 2, 4, 18 and 3%. Thus, the vast majority of Ero is partitioned to relative translation, with HF vibration coming in a distant second.

Calculating a single trajectory (ST) initiated at the TS, with a very small reaction coordinate translational energy and no energy in the TS's vibration (including ZPE) and rotation degrees of freedom, gives semi-quantitative values for partitioning of Ero to product energies. This model was used to study the effect of mass on energy partitioning for ‘C2H5F → HF + C2H4’ dissociation, where the MP2/6-31G* PES for C2H5F was used but masses of some of the atoms were changed for the ST direct dynamics simulations [27]. For the ST, 1.0 kcal mol−1 was added to reaction coordinate translation and the results of the simulations are given in table 2. Only changing the mass of F to Cl has a negligible effect on the product energy partitioning. However, changing the masses of the substituent H atoms to the mass of Cl has a dramatic effect. For the CHCl2CH2Cl model, where −CH2Cl contains the dissociating Cl atom, partitioning of energy to relative translation is dramatically lowered to 44%, with 14% and 26% partitioned to CCl2CH2 vibration and rotation, respectively. Changing the model to CH3CCl3 from CHCl2CH2Cl decreases and increases, respectively, energy partitioning to the olefin's rotation and vibration, with only a small change in the partitioning to relative translation. For CHCl2CCl3, with all of the substituent atoms Cl except one, the energy releases to relative translation and C2Cl4 vibration are nearly equivalent, while the release to C2Cl4 rotation is essentially zero.

Table 2.

Product energy partitioning for single trajectories using the MP2/6-31G* C2H5F → C2H4 + HF PESa.

molecule rel trans vibb rotb HX vib HX rot
C2H5F 74.9 6.8 1.5 14.4 2.4
C2H5Cl 75.6 5.9 2.4 15.9 0.2
CHCl2CH2Cl 43.9 14.2 26.0 15.8 0.1
CH3CCl3 47.3 24.4 11.5 15.7 1.1
CHCl2CCl3 39.7 38.1 0.2 16.1 5.9

aThe single trajectories were initiated at the saddlepoint with only 1.0 kcal mol−1 in reaction coordinate translation, Et.

bVib and rot are the vibration and rotation energies of the ethylene product, respectively.

The position of the centre of mass of the dissociating molecule affects partitioning of energy to olefin rotation for CHCl2CH2Cl, CH3Cl3 and CHCl2CCl3. Of these three molecules the largest percentage partitioning to olefin rotation is for CHCl2CH2Cl. For this molecule, the dissociation occurs by the H-atom of CHCl2 forming HCl with the Cl-atom of CH2Cl and then the HCl product recoils from the C-atom of the CH2 group of the product olefin. The centre of mass of the product olefin is located near the Cl atoms of the CCl2 group and the above recoil results in substantial rotation about this centre of mass, which is distant from the recoil site. For CH3CCl3, the product olefin's centre of mass is near the Cl atoms of the CCl2 group, but also near the recoil site (i.e. the C-atom of this group) and as a result energy transfer to olefin rotation is much smaller. The CCl2CCl2 olefin product is symmetric, for CHCl2CCl3 dissociation, with heavy atoms at both of its ends and a very large moment of inertia. It receives negligible rotational energy.

There are also significant mass effects for vibrational excitation of the olefin product. For C2H5F dissociation, only 6.8% of the product energy goes to olefin (i.e. C2H4) vibration, while for the CCl2CCl2 olefin product it is 38.1%. Analyses of the trajectories for CHCl2CCl3 → HCl + CCl2CCl2 dissociation show that two-thirds to three-fourths of the vibrational energy in the CCl2CCl2 product resides in the out-of-plane CCl2 wag-bend motions. There are two significant factors for this large wag-bend excitation and both are related to the mass of the Cl-atoms. As HCl dissociates from CHCl2CCl3, there is very little change in the CCl2 wag angles from their sp3 values for the dissociating molecule. HCl dissociates rapidly and there is insufficient time for the heavy Cl-atoms to move appreciably. A Franck–Condon type model [3133] approximates the resulting deposition of vibrational energy in the CCl2 wag-bends.

The other factor arises from the repulsive interaction between the Cl-atom of HCl and the C-atom from which Cl dissociates. If the C-atom is attached to two Cl-atoms, as HCl pushes off the C-atom, the Cl-atoms bonded to it move very little because of their substantial mass and the CCl2 wag-bend receives kinetic energy from the C-atom's motion. Another rather small Franck–Condon factor is also a mass effect. The Cl-atoms restrict the motion of the C-atoms, resulting in some vibrational excitation in the C=C stretch.

Of the three chlorinated model olefins, the energy partitioning to olefin vibration is 14.2, 24.4 and 38.1% for CHCl2CH2Cl, CH3Cl3 and CHCl2CCl3, respectively. These percentages are consistent with the factors discussed above. For CHCl2CH2Cl dissociation, the important factor for vibrational excitation of the CCl2CH2 product is the Franck–Condon excitation of the CCl2 wag-bend. For CH3Cl3 dissociation, there is this wag-bend excitation as well as the repulsive interaction between the dissociating Cl-atom and the C-atom of the CCl2 group. For CHCl2CCl3 dissociation, there are three factors: this repulsive interaction and Franck–Condon excitations of the two CCl2 wag-bends.

Exit-channel potential energy minimum. Many PESs have a potential energy minimum (or minima) in the exit channel connecting the TS and products. The statistical model for the exit-channel dynamics assumes the trajectories become trapped in these minima forming intermediates which, after rapid and complete intramolecular vibrational energy redistribution (IVR), dissociate to products with RRKM kinetics. If there is no reverse barrier for association of the reaction products to form the intermediate, the statistical model assumes there is a statistical distribution of energy for the reaction products as specified by PST [1215] or a modified PST [16,17]. However, this model may be inaccurate and the assumption of RRKM dynamics for the intermediate is then not correct.

An illustration of this type of non-statistical dynamics is found for the Cl + CH3Br → ClCH3 + Br SN2 reaction, which has a ClCH3‐‐‐Br complex in the exit channel [10,34]. If this complex is formed and its dynamics statistical, the ClCH3 + Br product energy partitioning should agree with the prediction of PST. In an experimental study of dissociation of the Cl‐‐‐CH3Br pre-reaction complex, Graul & Bowers [35,36] measured the relative translational energies between the ClCH3 + Br reaction products and found them to be much smaller than the prediction of PST. A chemical dynamics simulation was performed to model these experiments by exciting the Cl‐‐‐CH3Br complex and studying its dissociation. The intramolecular and unimolecular dynamics of this complex, and of the ClCH3‐‐‐Br exit-channel complex, were studied and found to be non-RRKM [10]. The post-TS dynamics after crossing the [Cl‐‐CH3‐‐Br] central barrier is decidedly non-RRKM. The average product relative translational energy found from the simulations is 0.8 kcal mol−1, in excellent agreement with the experimental value of 0.7 ± 0.2 kcal mol−1 and considerably different than the PST value of 1.6 kcal mol−1.

(iii). Avoiding exit-channel potential energy minima

Though there may be deep potential energy minima in the exit channel for chemical reactions, in the multi-dimensional post-TS dynamics the trajectories may avoid, i.e. not find, these minima. These dynamics are found for the F + CH3OH and F + CH3OOH reactions [37,38]. Energies and geometries for the potential energy curve of the F- + CH3OH SN2 reaction are shown in figure 3. The statistical model assumes the post-TS dynamics are indirect with the system temporarily trapped in the CH3OH‐‐‐F minimum. To test this model, direct dynamics trajectories were initiated at the central barrier TS, with conditions chosen from a 300 K Boltzmann distribution. Two reaction pathways were found for the trajectories, one direct and the other indirect. The vast majority, approximately 90%, follow a direct reaction pathway with departure of the F- ion approximately along the O-C‐‐‐F collinear axis, avoiding the potential energy minimum for the CH3OH‐‐‐F intermediate. The remaining small fraction of trajectories initially follow the direct pathway, but they do not have sufficient CH3OH + F relative translational energy to dissociate and are drawn into the CH3OH‐‐‐F minimum. As the system moves off the central barrier, it is propelled towards products by the potential energy release and does not move towards the potential energy well to form the CH3OH‐‐‐F intermediate.

Figure 3.

Figure 3.

Potential energy along the IRC for OH + CH3F → CH3OH + F. s is the distance along the IRC. The potential energy and IRC were calculated at the MP2/6-311G* level of theory. Adapted from [37].

Similar dynamics in which a deep exit-channel minimum is avoided is found for the F + CH3OOH reaction [38]. In experimental studies of this reaction [39], it was found that the most exothermic products HF + CH2(OH)O, and products predicted by the potential energy curve for the reaction (figure 4), are not formed. Instead, their experiments are consistent with formation of the much higher energy products HF + CH2O + OH by an ECO2 mechanism. To provide an atomic-level understanding of the experimental results, a direct dynamics simulation was performed [38]. Trajectories were initiated for the F + CH3OOH reactants with initial conditions to model the 300 K experiments. The post-TS dynamics for the trajectories, after passing TS1, did not lead to the deep potential energy of the CH2(OH)2‐‐‐F intermediate and instead directly formed the HF + CH2O + OH products by the proposed ECO2 mechanism. From the trajectories, the branching between the HF + CH2O + OH and HF + CH3OO product channels is predicted to be 0.98 ± 0.01 and 0.02 ± 0.01, fractions in qualitative agreement with the experimental estimates of approximately 85 and 10% for these two channels. The trajectory estimate of only approximately 2% branching to the HF + CH3OO products is not inconsistent with the experiments, based on a difficult numerical analysis. The trajectory total reaction rate constant for F + CH3OOH is (1.70 ± 0.7) × 10−9 cm3 molecule-s−1 and in excellent agreement with the experimental value of 1.23 × 10−9 cm3 molecule-s−1. As an additional test of the post-TS dynamics, trajectories initiated at TS1 with the energy of the reactants go directly to HF + CH2O + OH and to not move to the deep potential energy minimum.

Figure 4.

Figure 4.

Energy diagram for the F + CH3OOH reaction at the B3LYP/6-311+G(d,p) level of theory. The energies shown are in kcal mol−1 and are relative to the F + CH3OOH reactant channel. Zero-point energies are not included. The percentages are for the trajectories remaining in these regions of the PES when the B3LYP/6-311+G** direct dynamics trajectories are concluded at 4 ps. Adapted from [38]. (Online version in colour.)

The non-statistical post-TS dynamics found for the F + CH3OOH reaction is illustrated by the two-dimensional potential energy diagram in figure 5 [38]. Q1 represents the concerted movement of HF and OH away from CH2O, and Q2 represents the in-plane rotation motion of CH2O to access the deep potential energy minimum. Shown by the black line in figure 5 is a representative trajectory, which ‘skirts’ the deep potential energy minimum of the CH2(OH)2‐‐‐F. From the dynamics, the concerted movement of HF and OH away from CH2O is much faster than the rotation motion of CH2O needed to access the potential energy minimum. Also, there is no barrier restricting the movement of HF and OH away from CH2O. The red line illustrates the IRC leading to the complex. The dynamics for the F + CH3OOH reaction are reminiscent of those above for the OH + CH3F → CH3OH + F reaction.

Figure 5.

Figure 5.

A two-dimensional contour diagram of the post-transition state PES for TS1. Q1 = Δr1 + Δr2, where r1 is the FH–C bond length and r2 is the O─OH bond length. Q2 = Δθ1 + Δθ2, where Δθ1 is the O–C–O angle and Δθ2 is the H–O–C angle; i.e. H is the hydrogen abstracted by F, and O is the oxygen attached to carbon. Q1 represents the concerted motion of HF and OH away from CH2O, and Q2 represents rotation of CH2O. The remaining coordinates were optimized for each Q1, Q2 point. Depicted on this contour diagram is the IRC (red line) and a representative trajectory (black line). Adapted from ref. [38].

As found from direct dynamics simulation, deep exit-channel potential energy minima are also avoided for the microsolvated ions OH(H2O) and F(H2O) reacting with CH3I [40,41]. The reaction dynamics are similar for these two ions and those for OH(H2O) are discussed here. The trajectories were initiated for the OH(H2O) + CH3I reactants so that all possible reaction pathways could be studied. A schematic potential energy profile for the reaction is shown in figure 6. The reaction has a potential energy minimum for the CH3OH‐‐‐I(H2O) post-reaction complex for the SN2 reaction. The statistical model assumes this complex is formed and then dissociates to the various SN2 products. This complex was not formed in any of the trajectories. The dominant dynamics were for H2O to dissociate from the system as OH(H2O) attached to CH3I, with the associated potential energy release, or dissociate as OH displaced I with the much greater potential energy release. For the SN2 product ions I and I(H2O), there is a much greater yield of I. For OH + CH3I collision energies in the range of 2.0–0.05 eV, the I(H2O)/I ratio varies from 0.011 to 0.021 from the experiments and 0.018–0.041 from the simulations. The post-TS dynamics for the OH(H2O) + CH3I reaction are not consistent with a statistical model.

Figure 6.

Figure 6.

Schematic energy profile for the OH(H2O) + CH3I → CH3OH + I + H2O SN2 reaction, and other pathways, at the DFT/B97-1/ECP/d level of theory. The energies shown are in kcal mol−1 and are relative to the OH(H2O) + CH3I reactants. Zero-point energies are included. Adapted from ref. [40].

(b). Unimolecular decomposition

The statistical model for unimolecular dissociation assumes that IVR is rapid and complete for a unimolecular reactant so that its kinetics is accurately described by RRKM theory [7]. For numerous organic reactions, there is non-statistical unimolecular dynamics [42]. Non-RRKM behaviour is found in both experimental and simulation studies of X‐‐‐CH3Y pre-reaction complexes for X + CH3Y → XCH3 + Y SN2 reactions [18,43]. The energy released, when the reactants X and CH3Y associate, is initially only in the three intermolecular modes of the complex. Energy is then transferred to the nine CH3Y intramolecular modes by slow IVR. For a trajectory simulation of Cl + CH3Cl association at 300 K, the number of surviving Cl‐‐‐CH3Cl complexes versus time, N(t)/N(0), is accurately fit by a sum of three exponentials with rate constants varying from 1.1 to 0.093 ps−1 [44]. RRKM theory predicts a single exponential. When the trajectory N(t)/N(0) is averaged over collisions [45], a lifetime for the complex is obtained which is in excellent agreement with experiment as shown in figure 7 [46,47]. The lifetime does not agree with RRKM theory [45,46].

Figure 7.

Figure 7.

Average lifetime of the Cl‐‐‐CH3Cl complex, extracted from the Cl + CH3Cl → Cl‐‐‐CH3Cl ternary rate constant [46]. Included are values determined by high-pressure mass spectrometry [47] and chemical dynamics simulations [45]. Adapted from [18,46].

The unimolecular dynamics of the I‐‐‐CH3I complex, formed by I + CH3I association, has been studied by time-resolved photoelectron spectroscopy [48]. The time dependence of the I‐‐‐CH3I complex's lifetime, N(t)/N(0), is multi-exponential as discussed above for Cl‐‐‐CH3Cl. The unimolecular dynamics of this I‐‐‐CH3I complex are very similar to those observed in the trajectory study, described above, of the Cl‐‐‐CH3Cl complex. This is expected, since both complexes are formed by ion–molecule association, and the I + CH3I and Cl + CH3Cl intermolecular potentials are very similar.

Non-RRKM dynamics and kinetics are found for the Cl + CH3Br → ClCH3 + Br SN2 reaction [4951]. For this reaction, the [Cl‐‐CH3‐‐Br] central barrier TS is not rate controlling, and the kinetics of the Cl‐‐‐CH3Br pre-reaction complex must be modelled to calculate the SN2 reaction rate constant versus temperature T or reactant relative translational energy Erel. For the statistical model, RRKM theory is used to model the complex's kinetics, and this approach does not give rate constants versus either T or Erel which agree with experiment. Experimental studies of the Cl‐‐‐CH3Br complex have shown that its dynamics are non-RRKM [52]. When this complex is excited by a low-power continuous-wave CO2 laser at 943 cm−1, it only decomposes to ClCH3 + Br. By contrast, RRKM theory predicts that this excited complex should also decompose to Cl + CH3Br. The mode excited by the 943 cm−1 radiation is an intramolecular CH3 rocking mode. That only the ClCH3 + Br-products are observed is consistent with slow IVR between the intramolecular and intermolecular modes of the Cl‐‐‐CH3Br complex. Decomposition to form Cl + CH3Br requires energy transfer from the initially excited CH3 rock intramolecular mode to the Cl‐‐‐C stretch intermolecular mode.

In addition to the above SN2 studies, there are other chemical dynamics simulations which have found extensive non-statistical dynamics for organic reactions [42]. Non-RRKM dynamics has been observed in both experiments [53] and simulations [54] for the isomerization of the acetone-enol cation CH3–COH–CH2+ to the acetone cation CH3–CO–CH3+, which then dissociates to CH3+ + CH3CO and to CH3CO+ + CH3. What is found is that the methyl group formed by this isomerization is 1.36 ± 0.15 times more likely to dissociate from the acetone cation, while RRKM theory predicts equal dissociation probabilities for the two −CH3 groups.

Non-RRKM and non-statistical dynamics were found in chemical dynamics simulations of the ozonolysis of vinyl ethers, the results of which explain experimental observations [55]. The unimolecular dynamics of the primary ozonide, formed by the 1,3-dipolar addition of ozone to the vinyl ether, is non-statistical. The unimolecular kinetics of the ozonide was determined both experimentally and computationally, versus the size of the alkyl group of the vinyl ether, and the effect of the alkyl group size was found to be much less than the prediction of RRKM theory. IVR for the primary ozonide is incomplete and does not involve all of its vibrational modes on the time scale of its unimolecular decomposition. These results are consistent with the non-RRKM behaviour found for the simulations of the O3 + propene unimolecular dynamics [9].

(c). Non-intrinsic reaction coordinate dynamics

The above PES bifurcation, F + CH3OH, F + CH3OOH and OH(H2O) and F(H2O) + CH3I dynamics [2025,37,38,40,41] are examples of non-IRC dynamics. The IRC for the PES bifurcation leads to the VRI point and not to reaction products. The IRC for the F + CH3OH reaction leads to the CH3OH‐‐‐F post-reaction complex and formation of this complex is unimportant in the reaction dynamics. For the F + CH3OOH reaction, the IRC leads to the CH2(OH)2‐‐‐F complex and none of the randomly sampled trajectories form this complex. Such non-IRC dynamics are also found for cyclopropyl radical ring opening [56], 1,2,6-heptatriene rearrangement [57], the heterolysis rearrangement of protonated pinacolyl alcohol [58] and in the H + HBr → H2 + Br and O(3P) + CH3 → H2 + H + CO reactions [59,60]. Non-IRC dynamics may be important for the ‘roaming mechanism’ [61], which is discussed below, and non-IRC H-atom roaming was found in earlier studies of H-atom transfer [62,63]. Non-IRC dynamics is an important property of many chemical reactions.

(d). Bimolecular reactions

(i). Non-statistical dynamics

The statistical model assumes that the chemical reactants are trapped in pre-reaction complexes on the PES. However, chemical dynamics simulations show that SN2 nucleophilic substitution reactions, such as Cl + CH3Cl → ClCH3 + Cl and Cl + CH3Br → ClCH3 + Br, often occur by direct mechanisms without forming either the pre-reaction or the post-reaction ion-dipole complexes [64,65]; e.g. Cl‐‐‐CH3Br and ClCH3‐‐‐Br. For the Cl + CH3Cl reaction, exciting the C–Cl stretch mode of CH3Cl enhances the direct reaction mechanism [66]. The dominant mechanism for the SN2 reaction may change from indirect to direct over a small change in the reactant relative translational energy Erel. For the Cl + CH3Br reaction, this crossover occurs at Erel of approximately 6 kcal mol−1 [65]. In comparison, the well depth for the Cl‐‐‐CH3Br entrance-channel complex is approximately 10 kcal mol−1. There is a similar change from a dominant indirect to direct mechanism for the Cl + CH3I SN2 reaction [67], whose potential energy curve is shown in figure 1. The well depth for the Cl‐‐‐CH3I complex is 11.6 kcal mol−1. For Erel of 4.6 kcal mol−1, the indirect mechanism contributes 83% and 70% of the SN2 reaction as found from MP2/ECP/d and BHandH/ECP/d direct dynamics simulations, respectively. At the higher Erel of 9.0 kcal mol−1, only 1% and 12% of the reaction is indirect, as found by these two respective direct dynamics methods.

Direct reaction mechanisms are also important for the microsolvated reaction OH(H2O) + CH3I [40], whose potential energy curve is shown in figure 6. At the low collision energy of 0.05 eV (1.2 kcal mol−1), the reactants do not have sufficient energy to form the proton transfer products and only the SN2 pathways are open. As found from a B97-1/ECP/d direct dynamics simulation, there is no trapping in a post-reaction complex and 23% of the reaction occurs via a direct mechanism. Seventy-seven per cent of the trajectories have an indirect mechanism and are temporarily trapped in the (H2O)HO‐‐‐HCH2I and/or CH2I‐‐‐H2O(H2O) pre-reaction complex, as assumed by the statistical model.

In contrast to the prediction of RRKM and transition state theories, chemical dynamics simulations show that central barrier recrossing is important for the Cl + CH3Cl and Cl + CH3Br SN2 reactions [68,69]. These dynamics were studied for the Cl + CH3Cl SN2 reaction, first with an analytic PES [68] and then by MP2/6-31G* direct dynamics [70]. Trajectories were initiated at the [Cl‐‐CH3‐‐Cl] central barrier with a 300 K Boltzmann distribution of energy as assumed by TST and then integrated in both the reverse and forward directions. The dynamics predicted by RRKM theory is that the trajectories form the ion−dipole complexes in moving off the barrier (i.e. Cl‐‐‐CH3Cl for reverse and ClCH3‐‐‐Cl for forward) and then these complexes dissociate to reactants or products. The statistical kinetics assumes that crossing the barrier is rate controlling with no recrossings. The trajectories for the analytic PES and direct dynamics were integrated for 40 and 3 ps, respectively. RRKM theory predicts that 100% and 50% of the complexes should dissociate, for these respective simulations. By contrast, none of the complexes dissociated and the trajectories were ‘dynamically trapped’ in a region of the phase space about the [Cl‐‐CH3‐‐Cl] central barrier with extensive central barrier recrossings [68,70].

To quantify the importance of the central barrier recrossings, the number of trajectories residing in the post-reaction ClCH3‐‐‐Cl complex when the trajectories terminated was divided by the number of Cl‐‐‐CH3Cl → ClCH3‐‐‐Cl forward barrier crossings. This ratio is 0.1 for the analytic PES dynamics and 0.2 for the direct dynamics. Because none of the trajectories actually reached the ClCH3 + Cl product asymptote, more recrossings are expected if the trajectories were integrated for a longer time, resulting in a smaller value for the above ratio which TST assumes to be unity.

Central barrier recrossings are much less important for the Cl + CD3Cl and Cl + C2H5Cl SN2 reactions [71,72]. Trajectories were initiated at the [Cl‐‐CD3‐‐Cl] central barrier [71] and integrated in both the forward and reverse directions for 3 ps as for the [Cl‐‐CH3‐‐Cl] central barrier trajectories. Central barrier recrossings are much less important for Cl + CD3Cl and the above reaction to barrier recrossing ratio is 0.7 when compared with 0.2 for Cl + CH3Cl reaction. This difference may result from the lower intramolecular vibrational frequencies for CD3Cl when compared with CH3Cl, giving rise to more couplings between the low- and high-frequency vibrational modes of the reactive system and less phase space trapping in the vicinity of the central barrier. For the Cl + C2H5Cl SN2 reaction, there is negligible recrossing of the central barrier [72]. The majority of the trajectories move off the central barrier to form the Cl‐‐‐C2H5Cl complex and undergo efficient IVR and statistical/RRKM unimolecular dynamics to form Cl + C2H5Cl. However, a small number of the trajectories move directly to products without forming a complex, a non-RRKM result. The statistical dynamics for the Cl‐‐‐C2H5Cl complex my result from the low-frequency C2H5 torsion and bending modes which couple with the complex's intermolecular modes and facilitate IVR.

(ii). Reactant reactive states

As discussed above for C2H5F → HF + C2H4 dissociation, in moving from the TS to products, the product energy states (i.e. translation, rotation and vibration) are not populated in a statistical manner. Similar dynamics, with time reversal, are expected in moving from the TS to reactants [73]. These dynamics will identify the probabilities for the different reactant states to form products. The reaction is not statistical in the sense that the reactant states do not have identical probabilities for forming products.

For triatomic A + BC reactions, it is well understood whether translation or vibration excitation of the reactants promote reaction, which depends upon the position of the TS on the PES; i.e. early for translation and late for vibration. These have been identified as the ‘Polanyi rules’ [74]. These rules are not easily transferable to polyatomic bimolecular reactions [7578]. The state-specific dynamics of these reactions have been considered and analytic procedures have been described for determining their state specificities [77,78].

Xu et al. performed a B3LYP/6-31G* direct dynamics simulation [79], similar to the direct dynamics for C2H5F dissociation, except the trajectories were directed towards the reactants. Their simulation determined which reactant vibrational states have a high probability of attaining the TS structure and forming products for 1,3-dipolar addition reactions of diazonium betaines to acetylene and ethylene. The trajectories showed that reactant translation supplies the largest amount of energy needed to reach the TS, but reaction cannot proceed without a large amount of vibrational excitation in the bending mode of the betaine (i.e. N2O, N3H and CH2N2).

Analyses and calculations for enzyme catalysis reactions have suggested that protein dynamics may be important for some of these reactions [80,81]. Specific vibrations may promote the reaction on a short time scale, while particular conformational motions may be important for longer time dynamics. It may be possible to investigate each of these possible dynamical attributes by chemical dynamics simulations initiated at the TS and then directed towards the reactants. It may be possible to perform these simulations by QM/MM direct dynamics and/or MM chemical dynamics.

If the different reactant states have dissimilar probabilities for accessing the TS and forming products, one may say that these dynamics are non-statistical. However, this does not lead to a violation of TST. The validity of TST requires that: (i) a Boltzmann distribution is maintained for the translational, rotational and vibrational energies of the reactants during the reaction and (ii) once a trajectory crosses the TS in the direction reactants → products, the trajectory does not ‘turn around’ and recross [1,82]. Thus, the reactants → products flux may be taken as the reaction rate. If the reaction is dominated by high probabilities for a small number of reactant states, maintenance of a Boltzmann distribution for these states may become an important matter. In this manner, strong non-statistical bimolecular reaction dynamics might lead to non-TST kinetics.

(e). Roaming dynamics

There may be important non-statistical dynamics as part of the roaming mechanism for chemical reactions [61]. The term ‘roaming’ was coined in 2004 as part of a detailed discussion of the dynamics for H2CO → H2 + CO dissociation [83]. Initial descriptions of H2CO dissociation identified two pathways: i.e. the elimination pathway via a three-centred TS to form H2 + CO and a bond dissociation pathway to form the radicals H + HCO (figure 8). From chemical dynamics simulations, and in interpreting experimental measurements, a roaming mechanism was identified in which the loosely bound H-atom for H‐‐‐HCO dissociation ‘roamed’ around HCO and ultimately abstracted the H-atom from HCO forming H2 + CO. For an H + HCO collision, this would be identified as a disproportionation reaction [84]. For H2CO dissociation, roaming clearly couples the molecular elimination and bond dissociation pathways [61].

Figure 8.

Figure 8.

Schematics of a roaming trajectory in the dissociation of H2CO that initially moves in the direction of the radical H + HCO products, but turns in the high-energy region of the PES towards the high-energy isomers HOCH and ultimately undergoes an abstraction reaction to form H2 + CO. Panel (a) shows the energy profile of the PES and a cartoon of the roaming pathway. Panel (b) is a perspective plot of the PES illustrating the ‘turn’ made from the radical pathway towards the isomers. Adapted from [61].

Roaming has now been reported for numerous reactions, as described in [83]. Extensive analyses of roaming have been reported for CH3CHO [8587] and NO3 [8890] dissociation and isomerization in CH3NO2 [9193]. Roaming is expected to be important for hydrocarbon dissociation [94] and for C3H8 → CH3 + C2H5 may lead to the possible disproportionation products CH4 + C2H4 and CH2 + C2H6. Roaming for this reaction has a weak CH3‐‐‐C2H5 van der Waals interaction with possible minima and TSs in the roaming region of the PES.

The nature of the roaming dynamics, i.e. whether statistical or non-statistical, has received considerable attention [61]. A statistical theory for the kinetics and dynamics of roaming reactions has been proposed [95]. For roaming in formaldehyde, a statistical model may propose a TS for H‐‐‐HCO dissociation entering the roaming region of the PES and then a TS for H-atom abstraction to form H2 + CO. This statistical model would assume rapid IVR within the roaming region and RRKM kinetics for leaving this region. A statistical-like kinetic model has been proposed for the formation of and reaction from the formaldehyde roaming region, and rate constants for this model were determined from trajectory data [96].

Mauguière et al. [9799] have used a two-dimensional model, developed from the formaldehyde ab initio potential of Bowman and co-workers [100], to understand roaming dynamics for formaldehyde. They find that a phase space structure analysis of the roaming region is needed to understand the roaming dynamics [99]. Periodic orbit dividing surfaces are needed to delineate the different dynamics within the roaming region [101,102]. From the analysis of the phase structure for their model, they conclude that the roaming dynamics of formaldehyde are non-statistical and TST is not applicable.

A possible shortcoming of the phase space structure analysis for the formaldehyde roaming dynamics is that the Hamiltonian is a two-dimensional model and does not include all three H‐‐‐HCO intermolecular degrees of freedom. However, it is possible that the addition of the third intermolecular degree of freedom will not significantly alter the non-statistical dynamics found for roaming in formaldehyde. The HCO intramolecular degrees of freedom are expected to be only weakly coupled to the H‐‐‐HCO intermolecular modes. On the other hand, overall rotation and vibrational/rotational coupling could conceivably be important for the roaming intermolecular dynamics. Houston et al. [96] have studied formaldehyde trajectories using a full-dimensional potential. Azimuthal rotation of the roaming hydrogen atom about the CO bond axis of HCO was found to be a principal feature of the roaming. Roaming dynamics of reactive systems with more atoms may have low-frequency intramolecular modes which would couple to the intermolecular modes. Additional phase space structure analyses for roaming dynamics are expected.

3. Summary and future studies

The simulation and experimental studies described above illustrate that there are multiple non-statistical dynamics often observed for chemical reactions. Though simulation and theoretical studies of model systems may provide significant insight in non-statistical reaction dynamics, there is a sense that a deeper and more relevant understanding of the non-statistical dynamics may be obtained when both simulation and experiment tackle a particular problem. Examples of this are the studies for roaming in formaldehyde [61,96] and the dynamics of SN2 reactions [103].

What would be particularly meaningful is the development of unifying theoretical models for non-statistical dynamics. For statistical reaction dynamics, the RRKM and TS theories unify disparate experimental systems and provide a means for their comparisons. To illustrate important issues for non-statistical dynamics, consider non-IRC dynamics which avoid exit-channel potential minima on PESs [37,38] and, thus, form non-IRC reaction products [38]. May a theoretical model be developed which would predict when and for which reactions such dynamics are expected? For example, it may be possible to develop a theoretical model which relates non-IRC dynamics with strong curvature along the IRC reaction path [38,104].

Many of the above simulations involve direct dynamics [105]. This is a general approach for studying atomistic reaction dynamics, without the need for developing an analytic PES. As a means to interpret experiments and understand their non-statistical dynamics, many more direct dynamics simulations are expected. However, it is important to ensure that the electronic structure theory used for the direct dynamics gives accurate atomistic dynamics for the chemical reaction under investigation.

Acknowledgements

The authors wish to thank Stephen Wiggins for important comments and suggestions, regarding the material presented in this paper.

Competing interests

We declare we have no competing interests.

Funding

The atomistic simulations for the Hase research group are supported by the National Science Foundation under grant no. CHE-1416428, the Robert A. Welch Foundation under grant no. D-0005 and the Air Force Office of Scientific Research under AFOSR Award no. FA9550-16-1-0133.

References

  • 1.Steinfeld JI, Francisco JS, Hase WL. 1999. Chemical kinetics and dynamics, 2nd edn Upper Saddle River, NJ: Prentice Hall. [Google Scholar]
  • 2.Miller WH. 1993. Beyond transition-state theory: a rigorous quantum theory of chemical reaction rates. Acc. Chem. Res. 26, 174–181. ( 10.1021/ar00028a007) [DOI] [Google Scholar]
  • 3.Miller WH. 1998. Quantum and semiclassical theory of chemical reaction rates. Faraday Disc. Chem. Soc. 110, 1–21. ( 10.1039/a805196h) [DOI] [Google Scholar]
  • 4.Marcus RA, Rice OK. 1951. The kinetics of the recombination of methyl radical and iodine atoms. J. Phys. Colloid Chem. 55, 894–908. ( 10.1021/j150489a013) [DOI] [Google Scholar]
  • 5.Marcus RA. 1952. Unimolecular dissociations and free radical recombination reactions. J. Chem. Phys. 20, 359–364. ( 10.1063/1.1700424) [DOI] [Google Scholar]
  • 6.Rosenstock HM, Wallenstein MB, Wahrhafting AL, Eyring H. 1952. Absolute rate theory for isolated systems and the mass spectra of polyatomic molecules. Proc. Natl Acad. Sci. USA 38, 667–678. ( 10.1073/pnas.38.8.667) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Baer T, Hase WL. 1996. Unimolecular reaction dynamics. Theory and experiments. New York, NY: Oxford University Press. [Google Scholar]
  • 8.Mikosch J, Trippel S, Eichhorn C, Otto R, Lourderaj U, Zhang JX, Hase WL, Weidemüller M, Wester R. 2008. Imaging nucleophilic substitution dynamics. Science 319, 183–186. ( 10.1126/science.1150238) [DOI] [PubMed] [Google Scholar]
  • 9.Vayner G, Addepalli SV, Song K, Hase WL. 2006. Post-transition state dynamics for propene ozonolysis: intramolecular and unimolecular dynamics of molozonide. J. Chem. Phys. 125, 014317 ( 10.1063/1.2206785) [DOI] [PubMed] [Google Scholar]
  • 10.Lourderaj U, Park K, Hase WL. 2008. Classical trajectory simulations of post-transition state dynamics. Int. Rev. Phys. Chem. 27, 361–403. ( 10.1080/01442350802045446) [DOI] [Google Scholar]
  • 11.Fukui K. 1970. A formulation of the reaction coordinate. J. Phys. Chem. 74, 4161–4163. ( 10.1021/j100717a029) [DOI] [Google Scholar]
  • 12.Pechukas P, Light JC. 1965. On detailed balancing and statistical theories of chemical kinetics. J. Chem. Phys. 42, 3281–3291. ( 10.1063/1.1696411) [DOI] [Google Scholar]
  • 13.Klots CE. 1971. Reformulation of the quasiequilibrium theory of ionic fragmentation. J. Phys. Chem. 75, 1526–1532. ( 10.1021/j100680a025) [DOI] [Google Scholar]
  • 14.Chesnavich WJ, Bowers MT. 1977. Statistical phase space theory of polyatomic systems: rigorous energy and angular momentum conservation in reactions involving symmetric polyatomic species. J. Chem. Phys. 66, 2306–2315. ( 10.1063/1.434292) [DOI] [Google Scholar]
  • 15.Peslherbe GH, Hase WL. 1994. A comparison of classical trajectory and statistical unimolecular rate theory calculations of Al3 decomposition. J. Chem. Phys. 101, 8535–8553. ( 10.1063/1.468114) [DOI] [Google Scholar]
  • 16.Quack M, Troe J. 1975. Complex formation in reactive and inelastic scattering: statistical adiabatic channel model of unimolecular processes. III. Ber. Bunsenges. Phys. Chem. 79, 170–183. ( 10.1002/bbpc.19750790211) [DOI] [Google Scholar]
  • 17.Klippenstein SJ, Marcus RA. 1989. Application of unimolecular rate theory for highly flexible transition states to the dissociation of CH2CO to CH2 and CO. J. Chem. Phys. 91, 2280–2292. ( 10.1063/1.457035) [DOI] [Google Scholar]
  • 18.Manikandan P, Zhang J, Hase WL. 2012. Chemical dynamics simulations of X + CH3Y → XCH3 + Y gas-phase SN2 nucleophilic substitution reactions. Nonstatistical dynamics and nontraditional reaction mechanisms. J. Phys. Chem. A 116, 3061–3080. ( 10.1021/jp211387c) [DOI] [PubMed] [Google Scholar]
  • 19.Sun L, Hase WL. 2004. Ab initio direct dynamics trajectory simulation of C2H5F → C2H4 + HF product energy partitioning. J. Chem. Phys. 121, 8831–8845. ( 10.1063/1.1799573) [DOI] [PubMed] [Google Scholar]
  • 20.Thomas JB, Waas JR, Harmata M, Singleton DA. 2008. Control elements in dynamically determined selectivity on a bifurcating surface. J. Am. Chem. Soc. 130, 14 544–14 555. ( 10.1021/ja802577v) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Siebert MR, Zhang J, Addepalli SV, Tantillo DJ, Hase WL. 2011. The need for enzymatic steering in abietic acid biosynthesis: gas-phase chemical dynamics simulations of carbocation rearrangements on a bifurcating potential energy surface. J. Am. Chem. Soc. 133, 8335–8343. ( 10.1021/ja201730y) [DOI] [PubMed] [Google Scholar]
  • 22.Siebert MR, Manikandan P, Sun R, Tantillo DJ, Hase WL. 2012. Gas-phase chemical dynamics simulations on the bifurcating pathway of the pimaradienyl cation rearrangement: role of enzymatic steering in abietic acid biosynthesis. J. Chem. Theory Comput. 8, 1212–1222. ( 10.1021/ct300037p) [DOI] [PubMed] [Google Scholar]
  • 23.Quapp W, Bofill J. 2012. Topography of cyclopropyl radical ring opening to allyl radical on the CASSCF(3,3) surface: valley-ridge inflection points by Newton trajectories. J. Math. Chem. 50, 2061–2085. ( 10.1007/s10910-012-9995-8) [DOI] [Google Scholar]
  • 24.Collins P, Carpenter BK, Ezra GS, Wiggins S. 2013. Nonstatistical dynamics on potentials exhibiting reaction path bifurcations and valley-ridge inflection points. J. Chem. Phys. 139, 154108 ( 10.1063/1.4825155) [DOI] [PubMed] [Google Scholar]
  • 25.Kramer ZC, Carpenter BK, Ezra GS, Wiggins S. 2015. Reaction path bifurcation in an electrocyclic reaction: ring-opening of the cyclopropyl radical. J. Phys. Chem. A 119, 6611–6630. ( 10.1021/acs.jpca.5b02834) [DOI] [PubMed] [Google Scholar]
  • 26.Dong E, Setser DW, Hase WL, Song K. 2006. Comparison of levels of electronic structure theory in direct dynamics simulations of C2H5F → HF + C2H4product energy partitioning. J. Phys. Chem. A 110, 1484–1490. ( 10.1021/jp052888p) [DOI] [PubMed] [Google Scholar]
  • 27.Sun L, Park K, Song K, Setser DW, Hase WL. 2006. Use of a single trajectory to study product energy partitioning in unimolecular dissociation: mass effects for halogenated alkanes. J. Chem. Phys. 124, 064313 ( 10.1063/1.2166236) [DOI] [PubMed] [Google Scholar]
  • 28.Arunan E, Wategaonkar SJ, Setser DW. 1991. HF/HCl vibrational and rotational distributions from three- and four-centered unimolecular elimination reactions. J. Phys. Chem. 95, 1539–1547. ( 10.1021/j100157a008) [DOI] [Google Scholar]
  • 29.Quick CR Jr, Wittig C. 1980. IR multiple photon dissociation of fluorinated ethanes and ethylenes: HF vibrational energy distributions. J. Chem. Phys. 72, 1694–1700. ( 10.1063/1.439280) [DOI] [Google Scholar]
  • 30.Zamir E, Levine RD. 1980. Energy disposal in unimolecular elimination reactions. Chem. Phys. 52, 253–268. ( 10.1016/0301-0104(80)85229-3) [DOI] [Google Scholar]
  • 31.Berry MJ. 1974. Chloroethylene photochemical lasers: vibrational energy content of the HCl molecular elimination products. J. Chem. Phys. 61, 3114–3143. ( 10.1063/1.1682468) [DOI] [Google Scholar]
  • 32.Zvijac DJ, Light JC. 1977. Reactions of large molecules proceeding through an intermediate complex. I. Theory. Chem. Phys. 21, 411–431. ( 10.1016/0301-0104(77)85196-3) [DOI] [Google Scholar]
  • 33.Schatz GC, Ross J. 1977. Franck-Condon factors in studies of dynamics of chemical reactions. I. General theory, application to collinear atom-diatom reactions. J. Chem. Phys. 66, 1021–1036. ( 10.1063/1.434059) [DOI] [Google Scholar]
  • 34.Wang H, Goldfield EM, Hase WL. 1997. Quantum dynamical study of the Cl- + CH3Br SN2 reaction. J. Chem. Soc. Faraday Trans. 93, 737–746. ( 10.1039/a606061g) [DOI] [Google Scholar]
  • 35.Graul ST, Bowers MT. 1991. The nonstatistical dissociation dynamics of Cl(CH3Br): evidence for vibrational excitation in the products of gas-phase SN2 reactions. J. Am. Chem. Soc. 113, 9696–9697. ( 10.1021/ja00025a058) [DOI] [Google Scholar]
  • 36.Graul ST, Bowers MT. 1994. Vibrational excitation in products of nucleophilic substitution: the dissociation of metastable X(CH3Y) in the gas phase. J. Am. Chem. Soc. 116, 3875–3883. ( 10.1021/ja00088a024) [DOI] [Google Scholar]
  • 37.Sun L, Song K, Hase WL. 2002. A SN2 reaction that avoids its deep potential energy minimum. Science 296, 875–878. ( 10.1126/science.1068053) [DOI] [PubMed] [Google Scholar]
  • 38.López JG, Vayner G, Lourderaj U, Addepalli SV, Kato S, deJong WA, Windus TL, Hase WL. 2007. A direct trajectory study of F + CH3OOH reactive collision reveals a major non-IRC reaction path. J. Am. Chem. Soc. 129, 9976–9985. ( 10.1021/ja0717360) [DOI] [PubMed] [Google Scholar]
  • 39.Blanksby SJ, Ellison GB, Bierbaum VM, Kato S. 2002. Direct evidence for base-mediated decomposition of alkyl hydroperoxides (ROOH) in the gas phase. J. Am. Chem. Soc. 124, 3196–3197. ( 10.1021/ja017658c) [DOI] [PubMed] [Google Scholar]
  • 40.Xie J, Otto R, Wester R, Hase WL. 2015. Chemical dynamics simulations of the monohydrated OH(H2O) + CH3I reaction. Atomic-level mechanisms and comparison with experiment. J. Chem. Phys. 142, 244308 ( 10.1063/1.4922451) [DOI] [PubMed] [Google Scholar]
  • 41.Zhang J, Yang L, Xie J, Hase WL. 2016. Microsolvated F(H2O) + CH3I SN2 reaction dynamics. Insight into the suppressed formation of solvated products. J. Phys. Chem. Lett. 7, 660–665. ( 10.1021/acs.jpclett.5b02780) [DOI] [PubMed] [Google Scholar]
  • 42.Carpenter BK. 2005. Nonstatistical dynamics in thermal reactions of polyatomic molecules. Ann. Rev. Phys. Chem. 56, 57–89. ( 10.1146/annurev.physchem.56.092503.141240) [DOI] [PubMed] [Google Scholar]
  • 43.Hase WL. 1994. Simulations of gas-phase chemical reactions: applications to SN2 nucleophilic substitution. Science 266, 998–1002. ( 10.1126/science.266.5187.998) [DOI] [PubMed] [Google Scholar]
  • 44.Vande Linde SR, Hase WL. 1990. Trajectory studies of SN2 nucleophilic substitution. I. Dynamics of Cl + CH3Cl reactive collisions. J. Chem. Phys. 93, 7962–7980. ( 10.1063/1.459326) [DOI] [Google Scholar]
  • 45.Peslherbe GH, Wang H, Hase WL. 1995. Unimolecular dynamics of Cl‐‐‐CH3Cl intermolecular complexes formed by Cl + CH3Cl association. J. Chem. Phys. 102, 5626–5635. ( 10.1063/1.469294) [DOI] [Google Scholar]
  • 46.Mikosch J, Otto R, Trippel S, Eichhorn C, Weidemüller M, Wester R. 2008. Inverse temperature dependent lifetimes of transient SN2 ion-dipole complexes. J. Phys. Chem. A 112, 10 448–10 452. ( 10.1021/jp804655k) [DOI] [PubMed] [Google Scholar]
  • 47.Li C, Ross P, Szulejko JE, McMahon TB. 1996. High-pressure mass spectrometric investigations of the potential energy surfaces of gas-phase SN2 reactions. J. Am. Chem. Soc. 118, 9360–9367. ( 10.1021/ja960565o) [DOI] [Google Scholar]
  • 48.Wester R, Bragg AE, Davis AV, Neumark DM. 2003. Time-resolved study of the symmetric SN2 reaction I + CH3I. J. Chem. Phys. 119, 10 032–10 039. ( 10.1063/1.1618220) [DOI] [Google Scholar]
  • 49.Viggiano AA, Morris RA, Paschkewitz JS, Paulson JF. 1992. Kinetics of the gas-phase reactions of Cl with CH3Br and CD3Br: experimental evidence for nonstatistical behavior? J. Am. Chem. Soc. 114, 10 477–10 482. ( 10.1021/ja00052a050) [DOI] [Google Scholar]
  • 50.Craig SL, Brauman JI. 1997. Phase-shifting acceleration of ions in an ion cyclotron resonance spectrometer: kinetic energy distribution and reaction dynamics. J. Phys. Chem. A 101, 4745–4752. ( 10.1021/jp970602d) [DOI] [Google Scholar]
  • 51.Wang H, Hase WL. 1995. 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. 117, 9347–9356. ( 10.1021/ja00141a029) [DOI] [Google Scholar]
  • 52.Tonner DS, McMahon TB. 2000. Non-statistical effects in the gas phase SN2 reaction. J. Am. Chem. Soc. 122, 8783–8784. ( 10.1021/ja000881%2B) [DOI] [Google Scholar]
  • 53.Osterheld TH, Brauman JI. 1993. Infrared multiple-photon dissociation of the acetone enol radical cation. Dependence of nonstatistical dissociation on internal energy. J. Am. Chem. Soc. 115, 10 311–10 316. ( 10.1021/ja00075a054) [DOI] [Google Scholar]
  • 54.Nummela JA, Carpenter BK. 2002. Nonstatistical dynamics in deep potential wells: a quasiclassical trajectory study of methyl loss from the acetone radical cation. J. Am. Chem. Soc. 124, 8512–8513. ( 10.1021/ja026230q) [DOI] [PubMed] [Google Scholar]
  • 55.Quijano LMM, Singleton DA. 2011. Competition between reaction and intramolecular energy redistribution in solution: observation and nature of nonstatistical dynamics in the ozonolysis of vinyl ethers. J. Am. Chem. Soc. 133, 13 824–13 827. ( 10.1021/ja2043497) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Mann DJ, Hase WL. 2002. Ab initio direct dynamics of cyclopropyl radical ring opening. J. Am. Chem. Soc. 124, 3208–3209. ( 10.1021/ja017343x) [DOI] [PubMed] [Google Scholar]
  • 57.Debbert SL, Carpenter BK, Hrovat DA, Borden WT. 2002. The iconoclastic dynamics of the 1,2,6-heptatriene rearrangement. J. Am. Chem. Soc. 124, 7896–7897. ( 10.1021/ja026232a) [DOI] [PubMed] [Google Scholar]
  • 58.Ammal SC, Yamataka H, Aida M, Dupuis M. 2003. Dynamics-driven reaction pathway in an intramolecular rearrangement. Science 299, 1555–1557. ( 10.1126/science.1079491) [DOI] [PubMed] [Google Scholar]
  • 59.Pomerantz AE, Camden JP, Chiou A, Ausfelder F, Chawla N, Hase WL, Zare RN. 2005. Reaction products with internal energy beyond the kinematic limit result from trajectories far from the minimum energy path: an example from H + HBr → H2 + Br. J. Am. Chem. Soc. 127, 16 368–16 369. ( 10.1021/ja055440a) [DOI] [PubMed] [Google Scholar]
  • 60.Marcy TP, Díaz RR, Heard D, Leone SR, Harding LB, Klippenstein SJ. 2001. Theoretical and experimental investigation of the dynamics of the production of CO from the CH3 + O and CD3 + O reactions. J. Phys. Chem. A 105, 8361–8369. ( 10.1021/jp010961f) [DOI] [Google Scholar]
  • 61.Bowman JM. 2014. Roaming. Mol. Phys. 112, 2516–2528. ( 10.1080/00268976.2014.897395) [DOI] [Google Scholar]
  • 62.ter Horst MA, Schatz GC, Harding LB. 1996. Potential energy surface and quasiclassical trajectory studies of the CN + H2 reaction. J. Chem. Phys. 105, 558–571. ( 10.1063/1.471909) [DOI] [Google Scholar]
  • 63.Troya D, González M, Wu G, Schatz GC. 2001. A quasiclassical trajectory study of the Cl + HCN → HCl + CN reaction dynamics. Microscopic reaction mechanism of the H(Cl) + HCN → H2(HCl) + CN reactions. J. Phys. Chem. A 105, 2285–2297. ( 10.1021/jp003371a) [DOI] [Google Scholar]
  • 64.Hase WL, Cho YJ. 1993. Trajectory studies of SN2 nucleophilic substitution. III. Dynamical stereochemistry and energy transfer pathways for the Cl + CH3Cl association and direct substitution reactions. J. Chem. Phys. 98, 8626–8639. ( 10.1063/1.464470) [DOI] [Google Scholar]
  • 65.Wang Y, Hase WL, Wang H. 2003. Trajectory studies of SN2 nucleophilic substitution. IX. Microscopic reaction pathways and kinetics for Cl + CH3Br. J. Chem. Phys. 118, 2688–2695. ( 10.1063/1.1535890) [DOI] [Google Scholar]
  • 66.Vande Linde SR, Hase WL. 1989. A direct mechanism for SN2 nucleophilic substitution enhanced by mode selective vibrational excitation. J. Am. Chem. Soc. 111, 2349–2351. ( 10.1021/ja00188a086) [DOI] [Google Scholar]
  • 67.Zhang J, Lourderaj U, Sun R, Mikosch J, Wester R, Hase WL. 2013. Simulation studies of the Cl + CH3I SN2 nucleophilic substitution reaction: comparison with ion imaging experiments. J. Chem. Phys. 138, 114309 ( 10.1063/1.4795495) [DOI] [PubMed] [Google Scholar]
  • 68.Cho YJ, Vande Linde SR, Zhu L, Hase WL. 1992. Trajectory studies of SN2 nucleophilic substitution. II. Nonstatistical central barrier recrossing in the Cl + CH3Cl system. J. Chem. Phys. 96, 8275–8287. ( 10.1063/1.462331) [DOI] [Google Scholar]
  • 69.Wang H, Peslherbe GH, Hase WL. 1994. Trajectory studies of SN2 nucleophilic substitution. 4. Intramolecular and unimolecular dynamics of the Cl‐‐‐CH3Br and ClCH3‐‐‐Br complexes. J. Am. Chem. Soc. 116, 9644–9651. ( 10.1021/ja00100a032) [DOI] [Google Scholar]
  • 70.Sun L, Hase WL, Song K. 2001. Trajectory studies of SN2 nucleophilic substitution. 8. Central barrier dynamics for gas phase Cl + CH3Cl. J. Am. Chem. Soc. 123, 5753–5756. ( 10.1021/ja004077z) [DOI] [PubMed] [Google Scholar]
  • 71.Cheon S, Song K, Hase WL. 2006. Central barrier recrossing dynamics of the Cl + CD3Cl SN2 reaction. J. Mol. Struct.: THEOCHEM 771, 27–31. ( 10.1016/j.theochem.2006.03.032) [DOI] [Google Scholar]
  • 72.Sun L, Chang E, Song K, Hase WL. 2004. Transition state dynamics and a QM/MM model for the Cl + C2H5Cl SN2 reaction. Can. J. Chem. 82, 891–899. ( 10.1139/v04-082) [DOI] [Google Scholar]
  • 73.Barnes GL, Hase WL. 2009. Transition state analysis. Bent out of shape. Nat. Chem. 1, 103–104. ( 10.1038/nchem.193) [DOI] [PubMed] [Google Scholar]
  • 74.Polanyi JC. 1972. Concepts in reaction dynamics. Acc. Chem. Res. 5, 161–168. ( 10.1021/ar50053a001) [DOI] [Google Scholar]
  • 75.Liu J, Song K, Hase WL, Anderson SL. 2004. Direct dynamics trajectory study of vibrational effects: can Polanyi rules be generalized to a polyatomic system? J. Am. Chem. Soc. 126, 8602–8603. ( 10.1021/ja048635b) [DOI] [PubMed] [Google Scholar]
  • 76.Czakó G, Bowman JM. 2011. Dynamics of the reaction of methane with chlorine atom on an accurate potential energy surface. Science 334, 343–346. ( 10.1126/science.1208514) [DOI] [PubMed] [Google Scholar]
  • 77.Jiang B, Guo H. 2013. Relative efficiency of vibrational vs. translational excitation in promoting atom-diatom reactivity: rigorous examination of Polanyi's rules and proposition of sudden vector projection (SVP) model. J. Chem. Phys. 138, 234104 ( 10.1063/1.4810007) [DOI] [PubMed] [Google Scholar]
  • 78.Guo H, Jiang B. 2014. The sudden vector projection method for reactivity: mode specificity and bond selectivity made simple. Acc. Chem. Res. 47, 3679–3685. ( 10.1021/ar500350f) [DOI] [PubMed] [Google Scholar]
  • 79.Xu L, Doubleday CE, Houk KN. 2009. Dynamics of 1,3-dipolar cycloaddition reactions of diazonium betaines to acetylene and ethylene: bending vibrations facilitate reaction. Angew. Chem. Int. Ed. 48, 2746–2748. ( 10.1002/anie.200805906) [DOI] [PubMed] [Google Scholar]
  • 80.Quaytman SL, Schwartz SD. 2007. Reaction coordinate of an enzymatic reaction revealed by transition path sampling. Proc. Natl Acad. Sci. USA 104, 12 253–12 258. ( 10.1073/pnas.0704304104) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Antoniou D, Schwartz SD. 2011. Protein dynamics and enzymatic chemical barrier passage. J. Phys. Chem. B 115, 15 147–15 158. ( 10.1021/jp207876k) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Smith IWM. 1980. Kinetics and dynamics of elementary gas reactions. Theory and experiments. London, UK: Butterworths. [Google Scholar]
  • 83.Townsend D, Lahankar SA, Lee SK, Chambreau SD, Suits AG, Zhang X, Rheinecker J, Harding LB, Bowman JM. 2004. The roaming atom: straying from the reaction path in formaldehyde decomposition. Science 306, 1158–1161. ( 10.1126/science.1104386) [DOI] [PubMed] [Google Scholar]
  • 84.Laidler KJ. 1987. Chemical kinetics, 3rd edn. New York, NY: Harper & Row. [Google Scholar]
  • 85.Houston PL, Kable SH. 2006. Photodissociation of acetaldehyde as a second example of the roaming mechanism. Proc. Natl Acad. Sci. USA 103, 16 079–16 082. ( 10.1073/pnas.0604441103) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Heazlewood BR, Jordan MJT, Kable SH, Selby TM, Osborn DL, Shepler BC, Braams BJ, Bowman JM. 2008. Roaming is the dominant mechanism for molecular products in acetaldehyde photodissociation. Proc. Natl Acad. Sci. USA 105, 12 719–12 724. ( 10.1073/pnas.0802769105) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Rubio-Lago L, Amaral GA, Arregui A, González-Vázquez J, Bañares L. 2012. Imaging the molecular channel in acetaldehyde photodissociation: roaming and transition state mechanisms. Phys. Chem. Chem. Phys. 14, 6067–6078. ( 10.1039/c2cp22231k) [DOI] [PubMed] [Google Scholar]
  • 88.Grubb MP, Warter ML, Suits AG, North SW. 2010. Evidence of roaming dynamics and multiple channels for molecular elimination in NO3 photolysis. J. Phys. Chem. Lett. 1, 2455–2458. ( 10.1021/jz1008888) [DOI] [Google Scholar]
  • 89.Grubb MP, Warter ML, Xiao H, Maeda S, Morokuma K, North SW. 2012. No straight path: roaming in both ground- and excited-state photolytic channels of NO3 → NO + O2. Science 335, 1075–1078. ( 10.1126/science.1216911) [DOI] [PubMed] [Google Scholar]
  • 90.Fu B, Bowman JM, Xiao H, Maeda S, Morokuma K. 2013. Quasiclassical trajectory studies of the photodissociation dynamics of NO3 from the D0 and D1 potential energy surfaces. J. Chem. Theory Comput. 9, 893–900. ( 10.1021/ct3009792) [DOI] [PubMed] [Google Scholar]
  • 91.Homayoon Z, Bowman JM. 2013. Quasiclassical trajectory study of CH3NO2 decomposition via roaming mediated isomerization using a global potential energy surface. J. Phys. Chem. A 117, 11 665–11 672. ( 10.1021/jp312076z) [DOI] [PubMed] [Google Scholar]
  • 92.Homayoon Z, Bowman JM, Dey A, Abeysekera C, Fernando R, Suits AG. 2013. Experimental and theoretical studies of roaming dynamics in the unimolecular dissociation CH3NO2 to CH3 + NO2. Z. Phys. Chem. 227, 1267–1280. ( 10.1524/zpch.2013.0409) [DOI] [Google Scholar]
  • 93.Dey A, Fernando R, Abeysekera C, Homayoon Z, Bowman JM, Suits AG. 2014. Photodissociation dynamics of nitromethane and methyl nitrite by infrared multiphoton dissociation imaging with quasiclassical trajectory calculations: signatures of the roaming pathway. J. Chem. Phys. 140, 054305 ( 10.1063/1.4862691) [DOI] [PubMed] [Google Scholar]
  • 94.Sivaramakrishnan R, Michael JV, Harding LB, Klippenstein SJ. 2012. Shock tube explorations of roaming radical mechanisms: the decompositions of isobutane and neopentane. J. Phys. Chem. A 116, 5981–5989. ( 10.1021/jp210959j) [DOI] [PubMed] [Google Scholar]
  • 95.Klippenstein SJ, Georgievskii Y, Harding LB. 2011. Statistical theory for the kinetics and dynamics of roaming reactions. J. Phys. Chem. A 115, 14 370–14 381. ( 10.1021/jp208347j) [DOI] [PubMed] [Google Scholar]
  • 96.Houston PL, Conte R, Bowman JM. 2016. Roaming under the microscope: trajectory study of formaldehyde dissociation. J. Phys. Chem. A 120, 5103–5114. ( 10.1021/acs.jpca.6b00488) [DOI] [PubMed] [Google Scholar]
  • 97.Mauguière FAL, Collins P, Ezra GS, Farantos SC, Wiggins S. 2014. Multiple transition states and roaming in ion-molecule reactions: a phase space perspective. Chem. Phys. Lett. 592, 282–287. ( 10.1016/j.cplett.2013.12.051) [DOI] [Google Scholar]
  • 98.Mauguière FAL, Collins P, Ezra GS, Farantos SC, Wiggins S. 2014. Roaming dynamics in ion-molecule reactions: phase space reaction pathways and geometrical interpretation. J. Chem. Phys. 140, 134112 ( 10.1063/1.4870060) [DOI] [PubMed] [Google Scholar]
  • 99.Mauguière FAL, Collins P, Kramer ZC, Carpenter BK, Ezra GS, Farantos SC, Wiggins S. 2015. Phase space structures explain hydrogen atom roaming in formaldehyde decomposition. J. Phys. Chem. Lett. 6, 4123–4128. ( 10.1021/acs.jpclett.5b01930) [DOI] [PubMed] [Google Scholar]
  • 100.Zhang X, Zou S, Harding LB, Bowman JM. 2004. A global ab initio potential energy surface for formaldehyde. J. Phys. Chem. A 108, 8980–8986. ( 10.1021/jp048339l) [DOI] [Google Scholar]
  • 101.Pechukas P, McLafferty FJ. 1973. On transition-state theory and the classical mechanics of collinear collisions. J. Chem. Phys. 58, 1622–1625. ( 10.1063/1.1679404) [DOI] [Google Scholar]
  • 102.Pechukas P, Pollak E. 1979. Classical transition state theory is exact if the transition state is unique. J. Chem. Phys. 71, 2062–2068. ( 10.1063/1.438575) [DOI] [Google Scholar]
  • 103.Xie J, Otto R, Mikosch J, Zhang J, Wester R, Hase WL. 2014. Identification of atomic-level mechanisms for gas-phase X + CH3Y SN2 reactions by combined experiments and simulations. Acc. Chem. Res. 47, 2960–2969. ( 10.1021/ar5001764) [DOI] [PubMed] [Google Scholar]
  • 104.Wang H, Hase WL. 1996. Reaction path Hamiltonian analysis of the dynamics for Cl + CH3Br → ClCH3 + Br SN2 nucleophilic substitution. Chem. Phys. 212, 247–258. ( 10.1016/S0301-0104(96)00209-1) [DOI] [Google Scholar]
  • 105.Paranjothy M, Sun R, Zhuang Y, Hase WL. 2013. Direct chemical dynamics simulations: coupling of classical and quasiclassical trajectories with electronic structure theory. WIREs Comput. Mol. Sci. 3, 296–316. ( 10.1002/wcms.1132) [DOI] [Google Scholar]

Articles from Philosophical transactions. Series A, Mathematical, physical, and engineering sciences are provided here courtesy of The Royal Society

RESOURCES