Provided by the author(s) and NUI Galway in accordance with publisher policies. Please cite the published
version when available.
Title
Author(s)
A theoretical study of cyclic ether formation reactions
Bugler, John; Power, Jennifer; Curran, Henry J.
Publication
Date
2016-06-09
Publication
Information
Bugler, John, Power, Jennifer, & Curran, Henry J. (2017). A
theoretical study of cyclic ether formation reactions.
Proceedings of the Combustion Institute, 36(1), 161-167. doi:
https://doi.org/10.1016/j.proci.2016.05.006
Publisher
Elsevier
Link to
publisher's
version
https://doi.org/10.1016/j.proci.2016.05.006
Item record
http://hdl.handle.net/10379/6649
DOI
http://dx.doi.org/10.1016/j.proci.2016.05.006
Downloaded 2020-07-16T01:43:37Z
Some rights reserved. For more information, please see the item record link above.
Title Page
A Theoretical Study of Cyclic Ether Formation Reactions
John Buglera,*, Jennifer Powera, Henry J. Currana
a
Combustion Chemistry Centre, National University of Ireland, Galway, Ireland
Corresponding author:
John Bugler,
School of Chemistry,
National University of Ireland, Galway,
Ireland.
Email: j.bugler1@nuigalway.ie
Colloquium: REACTION KINETICS
1
A Theoretical Study of Cyclic Ether Formation Reactions
John Buglera,*, Jennifer Powera, Henry J. Currana
a
Combustion Chemistry Centre, National University of Ireland, Galway, Ireland
Abstract
Cyclisation reactions of hydroperoxyl-alkyl radicals forming cyclic ethers and hydroxyl radicals play
an important role in low temperature oxidation chemistry. These reactions contribute to the
competition between radical chain propagation and chain branching reaction pathways which
dominate the reactivity of alkanes at temperatures where negative temperature coefficient (NTC)
behaviour is often observed. The current study presents quantum chemically derived high-pressure
limit rate coefficients for all cyclisation reactions leading to cyclic ether formation in alkanes ranging
in size from C2 to C5. Ro-vibrational properties of each stationary point were determined at the M062X/6-311++G(d,p) level of theory. Coupled cluster (CCSD(T)) and Møller-Plesset perturbation
theory (MP2) methods were employed with various basis sets and complete basis set extrapolation
techniques to compute the energies of the resulting geometries. These methods, combined with
canonical transition state theory, have been used to determine 43 rate coefficients, with enough
structural diversity within the reactions to allow for their application to larger species for which the
use of the levels of theory employed herein would be computationally expensive. The validity of an
alternative, and computationally less expensive, technique to approximate the complete basis set
limit energies is also discussed, together with implications of this work for combustion modelling.
2
1. Introduction
In recent years, there has been a proliferation of systematic theoretical studies concerning reaction
pathways of importance in the low-temperature oxidation of alkanes [1–8]. The rapid increase in the
number of studies of this kind is due, in part, to more readily available computational resources.
This, coupled with the computationally inexpensive yet relatively accurate compound methods [9–
15], as implemented in the Gaussian software packages [16], has aided the investigation of large
arrays of reactions for which little or no experimental data are currently available. Several studies
have determined high-pressure limit rate coefficients for large numbers of reactions within the
important low-temperature reaction classes. Significant success has been achieved in chemical
kinetic modelling of alkane oxidation systems by utilising these values [17–19]. Despite these
successes, it has been highlighted that refinement of some important kinetic parameters is still
necessary for further progress [17]. Particular disparity is seen amongst literature values for the title
reactions. As is the case with many of the important uni-molecular reactions within the lowtemperature oxidation pathways, most theoretical rate coefficients for the reactions of interest in this
study have previously been derived using the CBS-QB3 compound method [9]. This is a popular
method, particularly when studying large quantities of species and/or reactions, due to its relative
speed and reliability [20]. Although relatively cost-effective, it has previously been highlighted as
being potentially biased towards the under-prediction of reaction barrier in some instances [21]. For
the reaction class of interest in this work, there may also be evidence of this based on investigations
by Villano et al. [7], where an average difference of approximately 2.4 kcal mol–1 in 0 K barriers is
observed between their values and those of DeSain et al. [22]. The barriers calculated by Villano et
al. using the CBS-QB3 method are consistently lower than those calculated for the same reactions by
DeSain et al., who used a combination of quadratic configuration interaction (QCISD(T)) and
Møller-Plesset perturbation theory (MP2) methods with varyingly sized basis sets (see [22] for more
details) to determine single point energies of stationary points characterised at the B3LYP/6-31G*
3
level of theory. In the instance of the cyclisation reaction of 4-hydroperoxyl-but-2-yl radical forming
2-methyloxetane and a hydroxyl radical, the 0 K barrier determined in both studies differs by 13.1
kcal mol–1! Further evidence of the under-prediction of reaction barriers for these reactions by the
CBS-QB3 method is observed in the recent study of Zhang et al. on the oxidation of n-hexane [19].
It was found that without significant modifications to the rate rules proposed by Villano et al. [7] for
this reaction class, model-predicted cyclic ether concentrations were too high when compared to
those measured in a jet-stirred reactor. Although the determination of accurate thermochemistry is
not the focus of this study, it is noteworthy that the CBS-QB3 method has recently been shown to
lack both accuracy and precision when deriving enthalpies of formation via the atomisation method
for a range of hydrocarbon and oxygenated species [20,23,24].
This study aims to provide high-fidelity rate coefficients for the reactions of interest through
utilisation of high-level quantum chemical methods. A comprehensive set of reactions is chosen in
order to allow application of the values derived in this work to similar reactions occurring in larger
molecules, for which the use of computational methods such as those employed here is currently
impractical.
2. Computational methods
2.1. Rate coefficient determination
All calculations have been performed using Gaussian 09 [16]. Geometries of minima and transition
state (TS) structures have been optimised using the M06-2X functional [25] with the 6-311++G(d,p)
basis set. Harmonic frequency analyses were employed at the same level of theory to verify the
nature of the stationary points, with a single imaginary frequency indicative of a first-order saddle
point on the potential energy surface (PES), corresponding to a TS structure. All frequencies were
scaled by 0.98, with zero point vibrational energies (ZPVEs) scaled by 0.97, as recommended for the
M06-2X functional by Zhao and Truhlar [25]. Intrinsic reaction coordinate (IRC) calculations [26]
4
were carried out with M06-2X/6-311++G(d,p) on each TS to ensure it was connected to the desired
reactants and products. Single point energy (SPE) calculations have been carried out for all C2H5O2
and C3H7O2 reactants and TSs using the coupled cluster (CCSD(T)) method and employing
relatively large basis sets (cc-pVTZ and cc-pVQZ [27]). The resulting energies were extrapolated to
the complete basis set (CBS) limit using the following formula [28,29]:
ECCSD(T)/CBS = ECCSD(T)/QZ + (ECCSD(T)/QZ – ECCSD(T)/TZ) 44 / 54 – 44
(1)
where TZ and QZ are abbreviations for cc-pVTZ and cc-pVQZ, respectively. For the C4H9O2 and
C5H11O2 species, the CCSD(T)/cc-pVQZ calculations were computationally prohibitive. For these
species, the CBS energies were estimated based on extrapolations of CCSD(T)/cc-pVDZ and
CCSD(T)/cc-pVTZ energies. The MP2 method was then used to correct for the difference in ccpVDTZ and cc-pVTQZ extrapolation energies. Here, cc-pVDTZ and cc-pVTQZ represent CBS
extrapolations based on cc-pVDZ and cc-pVTZ, and cc-pVTZ and cc-pVQZ energies, respectively.
The final energies were calculated using the following formula:
ECCSD(T)/CBS = ECCSD(T)/TZ + (ECCSD(T)/TZ – ECCSD(T)/DZ) 34 / 44 – 34 + EMP2/QZ + (EMP2/QZ – EMP2/TZ) 44 /
54 – 44 – EMP2/TZ – (EMP2/TZ – EMP2/DZ) 34 / 44 – 34
(2)
Similar approaches have previously been used to approximate “higher-level” SPEs [30–32]. The
validity of the approximation is investigated in this study by comparing the CCSD(T)/cc-pVTQZ
energies of the C2H5O2 and C3H7O2 species with those determined using Eq. (2). This is further
discussed in Section 3.1.
The T1 diagnostic [33] for all reactant species is ≤0.013, indicating that the use of single-reference
methods to describe the wave function is appropriate. T1 values for the TSs range from 0.031 to
0.040. While T1 values greater than 0.03 (for radicals) may be cause for concern [34], none of the
TSs have an unusually high value, with only six of the forty three complexes having a value greater
5
than 0.035. Nevertheless, if lower uncertainties are required for the TS energies, multi-reference
calculations are recommended.
Relaxed PES scans were carried out for internal rotations corresponding to low frequency torsional
modes in 10 degree increments as a function of dihedral angle with M06-2X/6-311++G(d,p).
Rotational constants were computed as a function of dihedral angle using the Lamm module of the
MultiWell program suite [35]. The resulting values were fitted to truncated Fourier series, and used
as input for 1-D hindered internal rotation approximations.
The Thermo application of MultiWell was used to compute high-pressure limit rate coefficients as a
function of temperature (298.15–2000 K) from canonical transition state theory [36]. Quantum
mechanical tunnelling has been accounted for via inclusion of 1-D tunnelling through an
unsymmetrical Eckart energy barrier [37]. The height of reaction barrier in the reverse direction is
required to account for tunnelling, yet we have not refined the SPEs of the reaction products past that
determined from the IRC analysis using M06-2X/6-311++G(d,p). This is due to the insensitivity of
the computed rate coefficients for reaction in the forward direction to the height of the barrier in the
reverse. Tests show that at 800 K a 10 kcal mol–1 variation in the reverse barrier results in a
difference of ~1% in rate coefficient in the forward direction. Therefore, we find it unnecessary to
determine the SPEs of the products to a higher accuracy. The resulting rate coefficients were fitted to
the following modified Arrhenius expression:
k = A (T/Tref)n exp(–E/RT)
(3)
where A is the A-factor, T is the temperature in units of Kelvin, Tref = 1 K, n is the temperature
exponent, and E is related to the activation energy (by Ea = E + nRT). This modified Arrhenius form
was adequate to represent the numerical data, with a maximum deviation of 14% between computed
and fitted rate coefficients. These expressions of the rate coefficients are listed in Table 1, Section
3.2.
6
2.2. Uncertainty
Zádor et al. [38] highlight an example of uncertainties in reaction barrier determinations for a
benchmark set of twenty H-atom abstraction reactions compiled by Lynch et al. [39]. Senosiain et al.
[40] tested a variety of methods against seventeen of these reactions in an attempt to quantify
uncertainty in reaction barrier determination for each method. Geometries were optimised using
either B3LYP [41] or MP2 [42] methods with the 6-311++G(d,p) basis set. SPE calculations were
carried out using B3LYP, MP2, QCISD(T), and CCSD(T) methods with augmented and nonaugmented cc-pVTZ and cc-pVQZ basis sets extrapolated to the CBS limit. Absolute error values are
lowest for the QCISD(T) and CCSD(T) methods, with >50% of the calculated barriers within 1 kcal
mol–1 of the benchmark values. The absolute errors for these cases appear to be largely independent
of the method used for geometry optimisation. The methods used here are quite similar to those used
by Senosiain et al., so their results may be useful for estimation of uncertainties in the barrier heights
presented here, although the are different from those compiled by Lynch et al. [39]. If it is assumed
that the uncertainties in barrier heights calculated in this study are normally distributed, with 50% of
the probability density function within 1 kcal mol–1 of the calculated value, we arrive at a 1σ
uncertainty of 1.5 kcal mol–1, which equates to a factor of ~2.6 uncertainty in rate constant at 800 K.
Estimating uncertainties in frequency factors is more difficult. The assumption that individual
contributions of hindered rotors are separable is likely to be adequate for the reactions of interest in
this study due to the lack of long-range interactions within the molecules. Interactions such as
hydrogen bonding tend to be more prevalent in molecules or complexes with multiple oxygenated
moieties, and leads to coupling of the internal rotors. This coupling, and the adequate treatment of
rotors when it occurs, has been discussed previously by Sharma et al. [1], and suitable methods were
applied to reactions of hydroperoxyl-alkyl-peroxyl radicals. Although the coupling of rotors is likely
to be significant in this study, neither is it likely that there is complete separability of rotors. On this
basis we estimate that uncertainties in frequency factors to be approximately a factor of 2.
7
3. Results and Discussion
3.1. Validity of CBS limit extrapolation approach
Barrier heights (E0 K + ZPVE) calculated using both Eq. (1) and Eq. (2) for the reactions of C2H5O2
and C3H7O2 species are compared, and are tabulated in the Supplementary Material, Table S1. It is
found that the barriers determined using coupled cluster and MP2 methods (denoted CC/MP2
hereafter) are consistently higher than those calculated using the coupled cluster (CC) method alone.
The difference in values is quite consistent, with Eq. 2 giving barriers which are higher by an
average of 0.46 kcal mol–1, with a 2σ dispersion of 0.09 kcal mol–1. The comparison set is small, but
with such a consistent offset it seems reasonable to lower the barriers calculated using Eq. (2) for all
of the reactions of C4H9O2 and C5H11O2 species by 0.46 kcal mol–1 from their CC/MP2 values. This
amount is within the uncertainty of the calculated barrier heights, but the aim is that this offset will
result in a more consistent set of values overall.
Commented [N1]: Not sure this is convincing.
3.2. Comparisons with the literature
Computed rate coefficients are presented in Table 1. Due to spatial constraints we provide
comparisons of these rate coefficients with literature values, as well as a detailed glossary of all
species listed in Table 1, as Supplementary Material. However, an account of the results is also given
here. As discussed in Section 1, Villano et al. [7] note that the barrier heights for these reactions
calculated by DeSain et al. [22] are an average of 2.4 kcal mol –1 higher than their own. The values
calculated in this study fall between those calculated in the two studies, with barrier heights an
average of 1.2 kcal mol–1 higher than those determined by Villano et al. [7]. This may provide yet
more evidence that CBS-QB3 tends to under-predict barrier heights.
8
Table 1. Rate coefficients calculated in this study.
Reaction
A (s–1)
n
E (cal mol–1)
Ċ2H4OOH1–2 ↔ C2H4O1–2 + ȮH
1.675 1007
1.40
10880.
Ċ3H6OOH1–2 ↔ C3H6O1–2 + ȮH
1.448 10
1.46
11850.
Ċ3H6OOH1–3 ↔ C3H6O1–3 + ȮH
05
7.556 10
1.56
18070.
Ċ3H6OOH2–1 ↔ C3H6O1–2 + ȮH
1.194 1008
1.26
11630.
Ċ4H8OOH1–2 ↔ C4H8O1–2 + ȮH
3.060 1007
1.41
11310.
Ċ4H8OOH1–3 ↔ C4H8O1–3 + ȮH
6.571 1004
1.79
16150.
Ċ4H8OOH1–4 ↔ C4H8O1–4 + ȮH
1.379 10
1.44
9920.
Ċ4H8OOH2–1 ↔ C4H8O1–2 + ȮH
08
5.058 10
1.11
11030.
Ċ4H8OOH2–3 ↔ C4H8O2–3anti + ȮH
8.979 1007
1.22
10260.
Ċ4H8OOH2–3 ↔ C4H8O2–3syn + ȮH
3.162 1008
1.04
9930.
Ċ4H8OOH2–4 ↔ C4H8O1–3 + ȮH
3.570 1006
1.36
16760.
Ċ4H8OOHI–I ↔ C4H8OI–I + ȮH
9.301 10
1.57
16670.
Ċ4H8OOHI–T ↔ C4H8OI–T + ȮH
2.642 10
1.35
10270.
Ċ4H8OOHT–I ↔ C4H8OI–T + ȮH
08
4.527 10
1.04
9930.
Ċ5H10OOH1–2 ↔ C5H10O1–2 + ȮH
4.667 1012
0.25
12840.
Ċ5H10OOH1–3 ↔ C5H10O1–3 + ȮH
1.275 1005
1.83
14460.
Ċ5H10OOH1–4 ↔ C5H10O1–4 + ȮH
4.728 1005
1.24
8130.
Ċ5H10OOH1–5 ↔ C5H10O1–5 + ȮH
2.314 10
1.31
8550.
Ċ5H10OOH2–1 ↔ C5H10O1–2 + ȮH
09
2.567 10
1.04
11340.
Ċ5H10OOH2–3 ↔ C5H10O2–3anti + ȮH
4.663 1008
1.09
9850.
Ċ5H10OOH2–3 ↔ C5H10O2–3syn + ȮH
7.464 1008
0.70
9270.
Ċ5H10OOH2–4 ↔ C5H10O2–4anti + ȮH
3.835 1006
1.26
14970.
Ċ5H10OOH2–4 ↔ C5H10O2–4syn + ȮH
1.392 10
1.49
15210.
Ċ5H10OOH2–5 ↔ C5H10O1–4 + ȮH
3.785 10
1.28
10220.
Ċ5H10OOH3–1 ↔ C5H10O1–3 + ȮH
06
2.398 10
1.52
17240.
Ċ5H10OOH3–2 ↔ C5H10O2–3anti + ȮH
2.814 1009
0.35
9860.
Ċ5H10OOH3–2 ↔ C5H10O2–3syn + ȮH
4.195 1009
0.71
10050.
Ċ5H10OOHA–A ↔ C5H10OA–A + ȮH
2.368 1005
1.77
16610.
07
05
05
07
04
06
05
Commented [N2]: Should we report these to three places of
decimals? 8.979 down below is pretty close to 8.98…
9
Ċ5H10OOHA–B ↔ C5H10OA–B + ȮH
2.440 1008
1.22
10420.
Ċ5H10OOHA–C ↔ C5H10OA–Canti + ȮH
1.158 1004
1.96
16160.
Ċ5H10OOHA–C ↔ C5H10OA–Csyn + ȮH
4.254 10
1.78
14740.
Ċ5H10OOHA–D ↔ C5H10OA–D + ȮH
05
5.363 10
1.27
9350.
Ċ5H10OOHB–A ↔ C5H10OA–B + ȮH
1.588 1009
0.85
9590.
Ċ5H10OOHB–C ↔ C5H10OB–C + ȮH
8.638 1008
0.85
8780.
Ċ5H10OOHB–D ↔ C5H10OB–D + ȮH
1.205 1007
1.22
16390.
Ċ5H10OOHC–A ↔ C5H10OA–Canti + ȮH
3.786 10
2.69
14900.
Ċ5H10OOHC–A ↔ C5H10OA–Csyn + ȮH
2.207 10
1.67
16240.
Ċ5H10OOHC–B ↔ C5H10OB–C + ȮH
10
1.191 10
0.77
9420.
Ċ5H10OOHC–D ↔ C5H10OC–D + ȮH
1.230 1009
1.01
10180.
Ċ5H10OOHD-A ↔ C5H10OA–D + ȮH
4.604 1005
1.30
9360.
Ċ5H10OOHD-B ↔ C5H10OB–D + ȮH
1.142 1005
1.65
13370.
Ċ5H10OOHD-C ↔ C5H10OC–D + ȮH
1.726 10
1.41
11350.
neoĊ5H10OOH ↔ neoC5H10O + ȮH
5.882 10
1.55
15990.
04
03
05
07
06
Figure 1. Comparison of literature values with this work for the cyclisation reaction of 3hydroperoxyl-prop-2-yl radical forming methyloxirane and a hydroxyl radical. Black: This
work, Red: Villano et al. [7], Blue: Miyoshi [3], Magenta: Wijaya et al. [43], Cyan: Cord et al.
[44], Wine: Chan et al. [45], Orange: Goldsmith et al. [31].
10
An example comparison of literature rate coefficients is shown in Fig. 1 for the cyclisation reaction
of 3-hydroperoxyl-prop-2-yl radical forming methyloxirane and a hydroxyl radical. This comparison
reflects the general trend seen when comparing literature rate coefficients with those calculated in
this study, in that those computed here tend towards the lower end of values which currently exist.
These lower rate coefficients may have been expected due to the findings of Zhang et al. [19], where
some of the rate rules suggested by Villano et al. [7] had to be lowered by approximately a factor of
4 in the temperature region where these reactions are most important (~700–900 K) in order to
improve model agreement with cyclic ether concentration profiles measured using a jet-stirred
reactor. This was achieved by lowering the A-factor of the Arrhenius expressions by a factor of 2,
and increasing the activation energy by 1 kcal mol–1.
3.3. Implications for combustion modelling
Figure 2 shows the effects of including the rate coefficients presented here for the reactions of
C5H11O2 to a recently published model describing oxidation of the pentane isomers [18]. Constant
volume and perfectly-stirred reactor simulations were run under some representative conditions in
which chemical kinetic models describing combustion processes are often validated. n-Pentane is
chosen as the representative fuel, and CHEMKIN-PRO [46] was used for the simulations. The closed
homogeneous batch reactor, and perfectly-stirred reactor modules within CHEMKIN-PRO were used
to simulate the ignition delay times and species concentration profiles, respectively. Ignition delay
simulations were run under stoichiometric fuel/‘air’ conditions (2.56% n-pentane, 20.46% O2,
76.98% N2) at 10 and 20 atm, and from 650–1400 K. The perfectly-stirred reactor simulations were
also run under stoichiometric conditions (1% n-pentane, 8% O2, 91% N2) at 1 atm, and from 500–
1100 K, at a residence time of 2 s. Also plotted are data presented in [18], Fig. 2 (a), as well as jetstirred reactor data yet to be published [47], Fig. 2 (b) and (c).
11
Figure 2. Model-simulated effects of rate coefficients presented in this study on n-pentane (a)
ignition delay times at 10 (black) and 20 atm (red), and perfectly-stirred reactor profiles of (b)
n-pentane, (c) 2-methyltetrahydrofuran (black), and 2,4-dimethyloxetane (red). Symbols
represent experimental data, dashed lines represent Model A, and solid lines Model B (see
Section 3.3 for details). The thicker lines in (a) represent simulations accounting for facility
effects.
The simulation results of two models (Model A and Model B) are shown. Model A is that presented
in [18], with the rate coefficients for C5 cyclic ether formation reactions taken from Villano et al. [7].
Model B is the same, but with the rate coefficients for the same reaction class replaced with those
pertaining to n-pentane from Table 1 in this study. It is shown that there is an increase in reactivity in
the NTC region in both sets of simulations, where the title reactions are known to be important, Fig.
2 (a) and (b). This reaction class is an important radical chain propagating one, and so this effect is as
expected given that the newly computed rate coefficients are lower than those from Villano et al. [7].
The effects seen are not big in terms of overall reactivity, but in Fig. 2 (c) the perfectly-stirred reactor
simulated concentration profiles of the two major cyclic ethers formed from n-pentane oxidation (2methyltetrahydrofuran and 2,4-dimethyloxetane) are shown, and a significant effect is seen. Two
peaks are observed in both concentration profiles at approximately 650 and 850 K, and factors of
~2–4 differences are seen in the simulated profiles at these temperatures. While model-predicted
mole fractions of 2-methyltetrahydrofuran have gone from slightly over-predicting the experimental
data to under-predicting it, those of 2,4-dimethyloxetane have improved considerably in terms of
12
agreement with experiment. While the graphs in Fig. 2 are mainly for illustrative purposes, it is seen
that the inclusion of the newly calculated rate coefficients into an existing model can bring about
significant changes and overall improvement in predicting cyclic ether concentrations. Model B
would require modifications in order to restore the accurate prediction of overall reactivity, but this
test provides insights into the modelling implications of using the rate coefficients presented in Table
1.
4. Conclusions
This paper presents a systematic and comprehensive study of the high-pressure limit kinetics of
cyclic ether formation reactions from hydroperoxyl-alkyl radicals. The rate coefficients are presented
and compared with those from the literature, and we find that those presented here are generally
lower than the existing values. Two different approaches are compared for the determination of
reaction barrier heights, and we validate a method which can approximate a “higher-level” answer at
a lower computational cost. The implications that these new rate coefficients may have for
combustion modelling are discussed, with results that are reasonably significant in terms of
mechanism predictions.
While this study presents values which are determined at a higher level of theory than other studies
for this reaction class, the modelling successes achieved by using values from these previous studies
cannot be understated. Several recent studies emanating from this research group and collaborators
have proven just how useful systematic studies of important reaction classes can be, even if the
accuracy of those values are not state-of-the-art. It is likely that these successes were possible due to
most of the rate coefficients for important reaction classes within low-temperature oxidation schemes
being calculated at the same level of theory (CBS-QB3). While the absolute accuracies of the values
are probably less than desirable, it may be the case that the relative values are more preferable,
resulting in favourable model predictions.
13
A more accurate determination of uncertainties in rate coefficients derived using different theoretical
methods would be extremely useful for chemical kinetic modellers. A benchmarking study of
different model chemistries, for instance, would go a long way in this regard. However, obtaining
suitable experimental data is likely difficult or currently impossible, so any such studies would have
to rely on comparisons with state-of-the-art theoretical calculations. Computationally less accurate
(and cheaper) methods with more accurate uncertainties may prove to be the most useful.
Acknowledgements
The authors wish to acknowledge the support of the Irish Research Council in funding this project
under project number EPSPG/2012/380, and also the provision of computational resources from the
Irish Centre for High-End Computing, ICHEC, under project number ngche026c. We also thank
Prof. John Simmie for helpful discussions and input on this work.
14
References
[1] S. Sharma, S. Raman, W. H. Green, J. Phys. Chem. A 114 (2010) 5689–5701.
[2] F. Zhang, T. S. Dibble, J. Phys. Chem. A 115 (2011) 655–663.
[3] A. Miyoshi, J. Phys. Chem. A 115 (2011) 3301–3325.
[4] J. Zádor, S. J. Klippenstein, J. A. Miller, J. Phys. Chem. A 115 (2011) 10218–10225.
[5] S. M. Villano, L. K. Huynh, H. -H. Carstensen, A. M. Dean, J. Phys. Chem. A 115 (2011) 13425–
13442.
[6] A. Miyoshi, Int. J. Chem. Kinet. 44 (2012) 59–74.
[7] S. M. Villano, L. K. Huynh, H. -H. Carstensen, A. M. Dean, J. Phys. Chem. A 116 (2012) 5068–
5089.
[8] S. M. Villano, H. -H. Carstensen, A. M. Dean, J. Phys. Chem. A 117 (2013) 6458–6473.
[9] J. A. Montgomery, M. J. Frisch, J. W. Ochterski, G. A. Petersson, J. Chem. Phys. 112 (2000)
6532–6542.
[10] J. W. Ochterski, G. A. Petersson, J. A. Montgomery, J. Chem. Phys. 104 (1996) 2598–2619.
[11] J. A. Pople, M. Head-Gordon, D. J. Fox, K. Raghavachari, L. A. Curtiss, J. Chem. Phys. 90
(1989) 5622–5629.
[12] L. A. Curtiss, C. Jones, G. W. Trucks, K. Raghavachari, J. A. Pople, J. Chem. Phys. 93 (1990)
2537–2545.
[13] L. A. Curtiss, K. Raghavachari, G. W. Trucks, J. A. Pople, J. Chem. Phys. 94 (1991) 7221–
7230.
15
[14] L. A. Curtiss, K. Raghavachari, P. C. Redfern, V. Rassolov, J. A. Pople, J. Chem. Phys. 109
(1998) 7764–7776.
[15] L. A. Curtiss, P. C. Redfern, K. Raghavachari, J. Chem. Phys. 126 (2007) 084108.
[16] M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G.
Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P.
Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R.
Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A.
Montgomery, J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N.
Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J.
Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo,
J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W.
Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J.
Dannenberg, S. Dapprich, A. D. Daniels, Ö. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski, D. J.
Fox, Gaussian 09, revision D.01; Gaussian, Inc., Wallingford, CT, 2009.
[17] J. Bugler, K. P. Somers, E. J. Silke, H. J. Curran, J. Phys. Chem. A 119 (2015) 7510–7527.
[18] J. Bugler, B. Marks, O. Mathieu, R. Archuleta, A. Camou, C. Grégoire, K. A. Heufer, E. L.
Petersen, H. J. Curran, Combust. Flame (2015) in press, doi:10.1016/j.combustflame.2015.09.014
[19] K. Zhang, C. Banyon, C. Togbé, P. Dagaut, J. Bugler, H. J. Curran, Combust. Flame 162 (2015)
4194–4207.
[20] J. M. Simmie, K. P. Somers, J. Phys. Chem. A 119 (2015) 7235–7246.
[21] J. Aguilera-Iparraguirre, H. J. Curran, W. Klopper, J. M. Simmie, J. Phys. Chem. A 112 (2008)
7047–7054.
16
[22] J. D. DeSain, C. A. Taatjes, J. A. Miller, S. J. Klippenstein, D. K. Hahn, Faraday Discuss. 119
(2001) 101−120.
[23] K. P. Somers, J. M. Simmie, J. Phys. Chem. A 119 (2015) 8922–8933.
[24] J. M. Simmie, J. Phys. Chem. A 119 (2015) 10511–10526.
[25] Y. Zhao, D. G. Truhlar, Theor. Chem. Account. 120 (2008) 215–241.
[26] H. P. Hratchian, H. B. Schlegel, J. Chem. Theory Comput. 1 (2005) 61–69.
[27] T. H. Dunning, J. Chem. Phys. 90 (1989) 1007–1023.
[28] J. M. L. Martin, Chem. Phys. Lett. 259 (1996) 669−678.
[29] D. Feller, D. A. Dixon, J. Chem. Phys. 115 (2001) 3484−3496.
[30] J. A. Miller, S. J. Klippenstein, J. Phys. Chem. A 107 (2003) 7783–7799.
[31] C. F. Goldsmith, W. H. Green, S. J. Klippenstein, J. Phys. Chem. A 116 (2012) 3325–3346.
[32] C. F. Goldsmith, G. R. Magoon, W. H. Green, J. Phys. Chem. A 116 (2012) 9033–9057.
[33] T. J. Lee, P. R. Taylor, Int. J. Quantum Chem.: Quantum Chemistry Symposium 23 (1989) 199–
207.
[34] S. J. Klippenstein, L. B. Harding, Proc. Combust. Inst. 32 (2009) 149–155.
[35] MultiWell-2014.1 Software, 2014, designed and maintained by J. R. Barker with contributors N.
F. Ortiz, J. M. Preses, L. L. Lohr, A. Maranzana, P. J. Stimac, T. L. Nguyen, and T. J. D. Kumar,
University of Michigan, Ann Arbor, MI, http://aoss.engin.umich.edu/multiwell/.
[36] H. Eyring, J. Chem. Phys. 3 (1935) 107–115.
[37] C. Eckart, Phys. Rev. 35 (1930) 1303–1309.
17
[38] J. Zádor, C. A. Taatjes, R. X. Fernandes, Prog. Energy Combust. Sci. 37 (2011) 371–421.
[39] B. J. Lynch, P. L. Fast, M. Harris, D. G. Truhlar, J. Phys. Chem. A 104 (2000) 4811–4815.
[40] J. P. Senosiain, S. J. Klippenstien, J. A. Miller, unpublished data.
[41] C. Lee, W. Yang, R. G. Parr, Phys. Rev. B 37 (1988) 785–789.
[42] C. Møller, M. S. Plesset, Phys. Rev. 46 (1934) 618–622.
[43] C. D. Wijaya, R. Sumathi, W. H. Green, J. Phys. Chem. A 107 (2003) 4908–4920.
[44] M. Cord, B. Sirjean, R. Fournet, A. Tomlin, M. Ruiz-Lopez, F. Battin-Leclerc, J. Phys. Chem. A
116 (2012) 6142–6158.
[45] W. -T. Chan, H. O. Pritchard, I. P. Hamilton, Phys. Chem. Chem. Phys. 16 (1999) 3715–3719.
[46] CHEMKIN-PRO 15101; Reaction Design: San Diego, 2010.
[47] ...
18