Вы находитесь на странице: 1из 14

ARTICLE

pubs.acs.org/JPCA

Tunneling in Hydrogen-Transfer Isomerization of n-Alkyl Radicals


Baptiste Sirjean,*,, Enoch Dames, Hai Wang,*, and Wing Tsang

Department of Aerospace and Mechanical Engineering, University of Southern California, Los Angeles, California 90089-1453,
United States

National Institute of Standards and Technologies, Gaithersburg, Maryland 20899, United States

bS Supporting Information
ABSTRACT: The role of quantum tunneling in hydrogen shift in
linear heptyl radicals is explored using multidimensional, smallcurvature tunneling method for the transmission coecients and a
potential energy surface computed at the CBS-QB3 level of theory.
Several one-dimensional approximations (Wigner, Skodje and
Truhlar, and Eckart methods) were compared to the multidimensional results. The Eckart method was found to be suciently
accurate in comparison to the small-curvature tunneling results for
a wide range of temperature, but this agreement is in fact fortuitous
and caused by error cancellations. High-pressure limit rate constants were calculated using the transition state theory with treatment of hindered rotations and Eckart transmission coecients for all
hydrogen-transfer isomerizations in n-pentyl to n-octyl radicals. Rate constants are found in good agreement with experimental kinetic data
available for n-pentyl and n-hexyl radicals. In the case of n-heptyl and n-octyl, our calculated rates agree well with limited experimentally
derived data. Several conclusions made in the experimental studies of Tsang et al. (Tsang, W.; McGivern, W. S.; Manion, J. A. Proc.
Combust. Inst. 2009, 32, 131138) are conrmed theoretically: older low-temperature experimental data, characterized by small preexponential factors and activation energies, can be reconciled with high-temperature data by taking into account tunneling; at low
temperatures, transmission coecients are substantially larger for H-atom transfers through a ve-membered ring transition state than
those with six-membered rings; channels with transition ring structures involving greater than 8 atoms can be neglected because of entropic
eects that inhibit such transitions. The set of computational kinetic rates were used to derive a general rate rule that explicitly accounts for
tunneling. The rate rule is shown to reproduce closely the theoretical rate constants.

1. INTRODUCTION
Understanding the thermal decomposition reactions of alkyl
radicals is critical to the development of combustion models of
hydrocarbon fuels. In n-alkane oxidation at high temperatures,
alkyl radicals are initially formed by CC or CH ssion in the
fuel molecule or H-abstraction reaction by a free radical. Small
alkyl radicals (Ce3) decompose mainly by CC -scission, as
the 12 and 13 hydrogen shifts can occur only through threeor four-membered ring transition state structures with large
energy barriers due to strain energies.1 For large alkyl radicals,
the thermal decomposition is complicated by internal hydrogen shifts.27 These isomerization processes occur through
cyclic transition state structures. Internal isomerization can occur
through ve-, six-, or seven-member ring transition state structures, which may be denoted as 14, 15, and 16 hydrogen
shifts, respectively. Some of these transition state structures are
associated with ring strain energies, while others are practically
unstrained. H-atom shifts are the predominant unimolecular
reactions under lower temperature conditions. As the temperature is increased, -scissions become increasingly important,
but H-atom shift remains critical to the dissociation of alkyl
radicals and especially the product distributions resulting from
the -scission processes. The combustion chemistry of olens
r 2011 American Chemical Society

and smaller alkyl radicals, formed in the decomposition process,


play a key role in the reactivity of the fuel and formation of
pollutants and polycyclic aromatic hydrocarbon (PAH) and
soot.812
The basic problem with studying alkyl isomerization processes
is that they involve large organic moieties (C5 or above). Except
at the lowest temperatures, -scission will always make contributions. Unless careful attention is paid to setting the reaction
conditions to minimize mechanistic artifacts, it is usually dicult
to derive truly quantitative kinetic parameters. In the case of single
pulse shock tube experiments, the isomerization rate constants
derived are all dependent on the initial -scission rate constants,
which, in turn, are derived from literature values. Thus, the
uncertainty in the -scission rate constants inevitably propagates
into the isomerization rate constants. For this type of tightly
coupled unimolecular reaction involving competitive reaction
paths, theoretical reaction kinetics is often necessary to interpret
the experimental data.
Received: September 28, 2011
Revised:
November 22, 2011
Published: November 30, 2011
319

dx.doi.org/10.1021/jp209360u | J. Phys. Chem. A 2012, 116, 319332

The Journal of Physical Chemistry A

ARTICLE

Figure 1. Arrhenius plots of (a) 14 hydrogen shift in n-pentyl radical, and (b) 15 hydrogen shift in n-hexyl radical.

Reactions involving the transfer of a hydrogen atom are inherently subject to quantum tunneling eect.13 The role of tunneling in n-alkyl radical isomerizations was recently highlighted by
Tsang and co-workers.47 They showed that tunneling must be
considered to reconcile older data obtained at low temperatures
with measurements made in their high-temperature single-pulse
shock tube. Low-temperature kinetic data have been subject to
controversies because of the unusually small pre-exponential factors
(A) that were dicult to rationalize with the thermochemical
kinetic theory.14 Direct measurements of the isomerization reaction
rates of alkyl radicals are dicult, and kinetic data were derived
usually from analyses of complex reaction mechanisms. Moreover,
most experiments were performed in the fallo region and/or
involved a chemically activated process. In such cases, the highpressure limit rate constant must be extrapolated from the Rice
RamspergerKasselMarcus (RRKM) theory and, for the latest
studies, from Master Equation (ME) analysis.
Among all n-alkyl radicals, the unimolecular reactions of
1-pentyl and 1-hexyl have been studied most extensively in the
past. Figure 1 presents selected literature rate constants. Endrenyi and Le Roy15 proposed the rst experimental rate constant
for the gas-phase 14 hydrogen migration in n-pentyl radicals:
1  C5 H11 / 2  C5 H11

based on a new mechanistic interpretation of the data of Endrenyi


and Le Roy and proposed:
k1 s1 5:0  1011 e10600=T

Similar low A factor values were observed from low-temperature


experiments on n-hexyl radicals. Watkins and Ostreko18 proposed
a rate expression for the 15 hydrogen migration in n-hexyl
radicals:
1  C6 H13 / 2  C6 H13

k2 s1 2:0  107 e4180=T

The small A factor has been a subject of debate in the


literature.19,20 In the late 1980s, Dobe et al.21 reported an experimental investigation of k2 for 300 e T e 453 K and pressures
from 100 to 200 Torr and proposed a rate expression that again
features a small A factor:
k2 s1 3:16  107 e5840=T

R1

Using the same experimental approach, these authors also proposed rate expressions for n-octyl radicals isomerization with
similar A factors. Marshall22 examined the thermal decomposition
of n-pentane in the temperature range of 737923 K and pressures below 200 Torr. The rate expression for the 14 hydrogen
shift of reaction R1 was derived from the distribution of major
products as

for temperatures 439 e T e 503 K at low pressures. Within the


framework of transition state theory, they concluded that the
unusual low pre-exponential factor can be explained only by
considering a quantum tunneling eect. Watkins16 derived a rate
expression from experiments performed 297 e T e 435 K and
pressures from 1 to 30 Torr:
k1 s1 3:3  108 e7600=T

R2

for temperatures from 352 to 405 K and a pressure of 46 Torr:

They reported:
k1 s1 1:4  107 e5440=T

k1 s1 9:1  1011 e11800=T

By considering earlier low-temperature measurements, he proposed that

k1 s1 1:2  1011 e10100=T

Watkins also noted the unusually low A factor of rate constant


and proposed that errors resulting from photochemical activation
could lead to an underestimation of the A factor. In a follow-up
paper,17 Watkins proposed a rate expression with a larger A factor

for 438 e T e 923 K. Imbert and Marshall23 followed a similar


approach to determine the high-pressure limit rate constant
15 hydrogen transfer in n-hexyl radicals by n-hexane pyrolysis
320

dx.doi.org/10.1021/jp209360u |J. Phys. Chem. A 2012, 116, 319332

The Journal of Physical Chemistry A

ARTICLE

and proposed:
k2 s1 3:16  1010 e8560=T

the framework of strain energy in the transition state structure.


Using B3LYP/ccpVDZ calculations and transition state theory
(TST) on prototypal reactions, Matheu et al.26 proposed rate
rules for 12, 13, 14, 15, and 16 hydrogen shifts in alkyl
radicals. Hayes and Burgess27 calculated the energy barriers of
hydrogen transfer in alkyl, allyilc, and oxoallylic radicals using
several composite methods and showed that an EvansPolanyi
correlation can be developed in the case of linear alkyl radicals.
Quantum tunneling eects were not considered in these studies.
Truong and co-workers28 studied the 1,4-intramolecular hydrogen migration in linear alkyl radicals using the class transition
state theory. They calculated the rate constant for the reference
reaction of n-C4H9 using canonical variational transition state
theory (CVTST) with the small curvature tunneling (SCT)
correction. It is important to note here that SCT approximation
allows for a multi-dimensional treatment of tunneling through
computationally expensive calculation of the Hessian for numerous points along the reaction path. In their study, the potential energy surface (PES) was computed at the CCSD(T)/
cc-pVDZ//BH&HLYP/cc-pVDZ level of theory. They found a
signicant tunneling contribution at temperatures below 1000 K.
Following their study on 14 hydrogen shift, they recently
proposed high-pressure limit rate expressions for 13 to 16
hydrogen migrations in linear alkyl radicals.29 They described the
isomerization reactions of n-propyl to n-hexyl radicals using the
same methodology. The rate constants for analogous reactions
were proposed on the basis of these prototype reactions within
the framework of reaction class TST. Again, the tunneling eect
was found to be less prominent above 1000 K. Additionally, they
compared transmission coecients obtained from multidimensional SCT tunneling with those obtained from the one-dimensional Eckart function and showed that the Eckart tunneling
eect is more pronounced than SCT tunneling below 400 K.
Jitariu et al.30 carried out direct dynamic calculations with
CVTST/SCT to study the decomposition and isomerization
pathways of n-pentyl radicals. They reported dual-level calculations at the PUMP2-SAC/6-311G**///AM1 level of theory.
Here, the triple slash (///) denotes interpolating optimized
corrections (IOC) in the VTST calculations.31 In this method,
the PES along the reaction path are corrected using higher-level
values at selected points. The Hessians, the potential energies,
and moment of inertia along the reaction path were computed
at the AM1 level of theory. UMP2/6-311G** geometries, vibrational frequencies, moments of inertia, and PUMP-SAC2/
6-311G** energies of the stationary points were then used to
scale the low-level AM1 results. This rather elaborate approach
led them to conclude that tunneling is pronounced at low
temperatures in the 14 hydrogen shift of 1-pentyl. At 1000
K, however, their proposed rate constant is about a factor 3 larger
than experimental values of Tsang et al.6,24 and Yamauchi et al.6,24
Zheng and Truhlar studied the 14 H shift in the 1-pentyl radical
and the 14 and 15 hydrogen transfers in the 1-hexyl radical
using CVTST/SCT theory.32 They calculated the PES with
several levels of density functional theory (DFT) (e.g., M062X/MG3S and M08-HX/cc-pVTZ+) and multilevel methods for
the stationary points. They used the interpolated variational
transition state theory by mapping (IVTST-M) in all of their
dynamics calculations. This method gives PES data (energies,
gradients, and Hessian) along the full reaction path based on a
limited number of points calculated near the saddle point. They
compared several one-dimensional approximations (Wigner, zero
curvature tunneling, and parabolic tunneling approximation) to

for 723 e T e 823 K. Miyoshi and co-workers derived the


expressions for k1 and k2 from an analysis of shock-tube experiments for 900 e T e 1400 K near the atmospheric pressure using
a complex reaction mechanism. They proposed expressions for
the high-pressure limit rate constants k from an RRKM analysis
of their experimental results along with earlier, low temperature
data. They included the tunneling eect via a Wigner correction
with the imaginary frequencies taken from HF/6-31G(d) calculations. They proposed:
24

k1, s1 4:9  108 T 0:846 e9830=T

and
k2, s1 6:7  107 T 0:823 e6570=T

10

for 350 e T e 1300 K. The somewhat strong and positive


temperature exponents point to an apparent upward curvature of
the Arrhenius form largely because of the tunneling eect being
considered in the rate calculation.
Miyoshi and co-workers25 reported the rst direct measurement of the rate coecients of 14 hydrogen shift in 1-pentyl
radical in the temperature range of 440520 K and pressures from 1 to 7 Torr. From an RRKM/ME analysis including
Eckart tunneling, they evaluated the high-pressure limit rate constant to be
k1, s1 1:9  1010 e8870=T

11

for 440 e T e 520 K and


k1, s1 2:4  103 T 2:324 e8180=T 9:1
 105 e5300=T

12

300 e T e 1300 K. The use of a bi-Arrhenius form is the result of


a large tunneling eect, leaving a high Arrhenius curvature.
More recently, Tsang and co-workers carried out a series of
systematic studies on the thermal decomposition of n-pentyl,6
n-hexyl,7 n-heptyl,4 and n-octyl5 radicals in single-pulsed shock
tube for temperatures from 850 to 1050 K and pressures between
1.5 and 7 bar. From experimentally determined product branching ratios, they deduced high-pressure limit rate expressions for
14, 15, and 16 hydrogen transfers using an RRKM/ME
analysis. In the case of n-octyl, they also explored a hydrogen
migration channel through the eight-membered ring transition
state structure, but found that such a transition is unimportant
compared to other reaction channels. Earlier, lower temperature
data on n-pentyl and n-hexyl were included in their theoretical
analysis considering Eckart tunneling. Notably, they proposed
the high-pressure limit rate constant:
k1, s1 1:0  1012 e11300=T

13

for 14 hydrogen shift in 1-pentyl from 850 to 1000 K, and


k2, s1 1:8  102 T 2:55 e5550=T

14

for 15 hydrogen migration in 1-hexyl for 500 e T e 1900 K.


Theoretically, Viskolcz et al.1 computed the energy barriers of
several isomerization reactions of linear alkyl radicals and the
barrier of the 14 hydrogen shift in 1-pentyl at the MP-SAC2//
UHF/6-31G* level of theory. They discussed their results within
321

dx.doi.org/10.1021/jp209360u |J. Phys. Chem. A 2012, 116, 319332

The Journal of Physical Chemistry A

ARTICLE

(ZPE) were quite close at the B3LYP/6-311G(2d,d,p) and CBSQB3 levels of theory (Table 1).
Hindered Internal Rotations. The description of isomerization channels of large aliphatic chains requires an accurate treatment of low-frequency internal rotations. Using the harmonic
oscillator (HO) approximation to describe these torsion modes
can lead to large errors in the partition function. Vansteenkiste
et al.38 calculated the thermodynamic properties of n-alkanes
using a quantum-mechanical treatment and showed that treating
hindered internal rotation is critical to obtain accurate entropies
and heat capacities values. As hydrogen transfers occur through
cyclic transition state structures, most of the internal rotations of
the initial reacting radicals are locked in the ring, and the rate
coefficients will be strongly dependent on the entropy variation.
As an example, Vansteenkiste et al. reported for n-heptane a
difference of 9.3 cal/(mol K) at 1000 K in entropy between
calculations using the assumption of harmonic oscillator (HO)
and a more precise quantum mechanical treatment. The HO
approximation underestimates the entropy change. In the present work, hindered internal rotations were treated using the
following procedure. First, the potentials of each internal rotation
of 1-pentyl, 2-pentyl, and 3-pentyl radicals were calculated at the
B3LYP/6-311G(2d,d,p) level of theory using a relaxed energy
scan. The energy barriers for these hindered internal rotors were
then calculated at the CBS-QB3 level of theory. The characteristics of the rotational potentials and the barriers of rotation
obtained were used to correct the HO partition function using
Pitzer and Gwinn tabulations.39 Rotational potential functions and
barriers for each internal rotor in n-pentyl were used for similar
internal rotors found in largest alkyl radicals. Reduced moments of
inertia for internal rotations were calculated for each species using
B3LYP/6-311G(2d,d,p) geometries with the method of Pitzer
implemented in ChemRate.40,41 For cyclic transition states, the
vibrational modes of the cyclic part of the molecular structure were
described within the HO approximation, and the lateral alkyl
group internal rotations were treated as hindered rotors (HR) via
relaxed scans (with cyclic bond lengths frozen) in the critical geometries for 1-pentyl to 2-pentyl and 1-hexyl to 3-hexyl. Parameters
for HR corrections for these two critical geometries were used for
all other similar internal rotors found in the saddle point geometries of larger alkyl radicals. Hindrance potentials and barrier
heights are given in the Supporting Information.
Transmission Coefficient. Within a canonical TST or VTST
framework, quantum tunneling is taken into account by a
temperature-dependent transmission coefficient k(T):

Table 1. Critical Energies (kcal/mol) Computed for H-Atom


Shifts in n-Heptyl Radicals at 0 K
B3LYP/
7ps (1-heptyl / 2-heptyl)
6ps (1-heptyl / 3-heptyl)

6-311G(2d,d,p)a

CBS-QB3a

G3MP2B327

14.1
14.6

14.7
15.0

15.8
16.1

5ps (1-heptyl / 4-heptyl)

21.3

22.0

23.5

5ss (2-heptyl / 3-heptyl)

23.2

22.8

23.9

This work.

their multidimensional SCT results. They noted that at low


temperatures the discrepancies among these methods are rather
large. An interesting point is that tunneling is more important in
the case of 14 hydrogen transfer than for 15 shift, in agreement with the work of Tsang and co-workers. Unfortunately, the
widely used Eckart method was excluded from their study, and
its ability to reproduce results of higher dimensional tunneling
treatments remains unclear.
In the present study, we intend to address several issues
regarding the role of tunneling in linear alkyl radical isomerizations. We compute the transmission coecients using n-heptyl as
the model system employing a multi dimensional treatment. The
results allow us to assess accuracy of the various one-dimensional
methods for the calculation of transmission coecients. Beyond
the n-heptyl radical, we use the rigid rotor harmonic oscillator
(RRHO) approximation with corrections for hindered internal
rotation (HR), the TST theory with a selected 1-D tunneling
method, and we systematically determine the rate coecients for
14 to 16 H-shifts in n-alkyl radicals ranging from n-pentyl to
n-octyl. Note that as the size of the linear alkyl chain increases,
most favorable isomerizations (ve-, six-, and seven-member
transition state structures) can involve hydrogen shift between
two secondary radicals (e.g., 2-heptyl to 3-heptyl). It is noteworthy to mention that there are almost no kinetic data available
for these types of processes. Finally, we discuss the role of
tunneling within the framework of correlations between structure and reactivity, and its impact on the uncertainty of rate
coecients derived from theoretical calculations.

2. COMPUTATIONAL DETAILS
Electronic Structure Calculations. Potential energy and
molecular properties of stationary points were calculated using
Gaussian 03 revision B.0333 and QChem version 3.1.34 For all
stationary and critical structures, geometry optimizations were
performed at the B3LYP/6-311G(2d,d,p) level of theory.35,36
Frequency calculations were performed at the same level of
theory for all optimized geometries to determine the nature of
stationary points. The composite CBS-QB3 method was applied
for all stationary geometries and transition states.37 The CBSQB3 model involves a five-step calculation starting with a
geometry optimization and a frequency calculation (scaled by a
factor 0.99) at the B3LYP/6-311G(2d,d,p) level of theory,
followed by single point energy calculations at the CCSD(T)/
6-31G(d0 ), MP4SDQ/cbsb4, and MP2/cbsb3 and a complete
basis set extrapolation to correct the total energy. Dynamic calculations along the reaction path were performed at the B3LYP/
6-311G(2d,d,p) level of theory. We found that for reactions
examined here, the energy barrier heights with zero-point energy

kT kTAeE=T

15

Within the framework of the RRKM theory and to account for


tunneling, Miller42 proposed one-dimensional tunneling probability P(E,J) in the sum of states N(E,J) of the transition state:
NE, J

n PE  Eqn , J

16

where nq is the nth vibrational energy level and J is the angular


momentum. The standard microcanonical rate expression is
calculated using N(E,J):
kE, J

NE, J
hFE, J

17

where h is the Planck constant and F(E,J) is the density of energy


states of the reactant. Note that k(T) can be calculated from P(E,J)
322

dx.doi.org/10.1021/jp209360u |J. Phys. Chem. A 2012, 116, 319332

The Journal of Physical Chemistry A

ARTICLE

by integrating over the Boltzmann-weighted energy distribution


and that k(T) also includes nonclassical reflection effects.
For a one-dimensional system, the transmission coecient
may be calculated by solving the Schr
odinger equation with an
appropriate potential function. In this approach, the potential
energy surface is tted with a function for which the tunneling
probabilities are known analytically. The simplest method is that
of Wigner43 who assumed a parabolic function for the PES
leading to the transmission coecient:
"
#2
1 hImq
kT 1
24 kB T

18

where q is the imaginary frequency corresponding to the


reaction coordinate, and kB is the Boltzmann constant. Skodje
and Truhlar44 proposed an improvement over the Wigner
method by treating the energy barrier as a truncated parabola.
The Eckart function is particularly useful as it provides realistic
potential function that has the correct asymptotic properties of
the one-dimensional PES. Moreover, the parameters of the
Eckart potential can be tted to reproduce the curvature at the
saddle point and the exothermicity of the reaction.45 This onedimensional method is widely used in the theoretical combustion
kinetics because they are rather inexpensive computationally yet
accurate enough for the temperatures of interest.
More sophisticated methods to calculate the transmission coecient have been proposed by Truhlar and co-workers. Particularly, they developed semiclassical methods to calculate transmission coecients that take into account the multidimensional
nature of tunneling; that is, tunneling paths deviate dramatically from the reaction path. These methods are computationally
expensive as they required detailed information (e.g., Hessians)
along the reaction path. For reactions with a path curvature assumed to be small, the small curvature tunneling (SCT) method
was proposed.46,47 It allows for corner-cutting approximately on
the concave side of the turning points of the vibrations transverse
to the minimum energy path. For larger curvatures, complications
arise as contributions from tunneling into excited states may have
to be considered. In that case, the large curvature tunneling (LCT)
method can be used to calculate transmission coecients.48
For the n-heptyl radicals, we carried out calculations of
transmission coecients with the SCT method. The minimum
energy paths (MEP) were determined at the B3LYP/6-311G
(2d,d,p) level of theory with direct dynamics calculations. MEP
calculations were performed in Cartesian coordinates with a step
size of 0.01 using the Euler steepest-descents integrator. This
step size was found to be suciently small to converge the
reaction path and the transmission probabilities in the range of
reaction coordinates ranging from 2.0 to +2.5 for 16 and
15 hydrogen shifts, and 2.8 to 3 and 2.5 to +3 for 14
and 25 hydrogen migrations, respectively. All of the SCT
transmission coecients reported here were calculated using
the POLYRATE 9.749 and GAUSSRATE 9.750 codes. LCT
calculations were also performed to determine if a large curvature
mechanism could be important for the range of reaction coordinate studied here. We do not expect to obtain accurate
transmission coecients from the LCT method as the reaction
coordinate ranges may be too small for a proper description of a
large curvature path, and it may be necessary to include tunneling
in excited states to obtain accurate results.

Figure 2. Potential energy of H-atom shift in the n-heptyl radicals


computed using the CBS-QB3 method at 0 K.

One-dimensional transmission coecients were also computed using Wigner and Skodje and Truhlar approximations
directly from CBS-QB3 results on stationary points of the PES.
The one-dimensional Eckart transmission coecients were calculated using ChemRate software where the characteristic length
of the Eckart function is obtained from forward and reverse
barrier heights (E1 and E1) at 0 K along with the imaginary
frequency (i) of the transition state using the equations reported
by Johnston and Heicklen.51 Note that in Chemrate, canonical
transmission coecients are calculated by integrating microcanonical transmission probabilities over the Boltzmann-weighted
energy distribution. Lennard-Jones parameters were taken from
the JetSurF version 1.0 transport database.52,53 Calculated characteristic lengths, as well as parameters E1, E1, and i, are given in
the Supporting Information.

3. RESULTS AND DISCUSSION


Potential Energy Surfaces. The potential energy for hydrogen shift in n-heptyl is presented in Figure 2. As the heptyl
radicals allow for secondary to secondary H shift, we will follow
the notation proposed by Hardwidge et al.19 that explicitly retains
this information. Within this framework, a particular hydrogen
transfer will be described as an iab process, where i is the ring size
of the cyclic transition state structure, a refers to a primary or
secondary radical (noted p or s) for a reactant, and b is p or s for a
product. Within this nomenclature, the 14 hydrogen transfer of
1-pentyl radical would be referred to as a 5ps isomerization.
It is known that transmission coecients are sensitive to the
barrier heights. Therefore, to be able to analyze the SCT transmission coecients obtained with molecular parameters obtained at the B3LYP/6-311G(2d,d,p) level of theory and the
Eckart k from CBS-QB3//B3LYP/6-311G(2d,d,p), the critical
energies computed at these two level of energies were compared.
From Table 1, it can be seen that the CBS-QB3 and DFT energy
barriers computed for heptyl are very close to each other with the
largest deviation being 0.7 kcal/mol. Consequently, comparisons
can be made directly between the one-dimensional and SCT
transmission coecients.
Comparisons are also made with the G3MP2B3 results of
Hayes and Burgess27 (Table 1). The CBS-QB3 critical energies
are found to be systematically lower than the G3MP2B3 values.
The discrepancies range from 1.1 kcal/mol (5ss) to 1.5 kcal/mol
(5 ps), with a mean absolute deviation of 1.2 kcal/mol. These
dierences are perhaps within the limits of the uncertainty of the
323

dx.doi.org/10.1021/jp209360u |J. Phys. Chem. A 2012, 116, 319332

The Journal of Physical Chemistry A

ARTICLE

300 and 400 K than the present results. In particular, their Eckart
transmission coefficients are larger than our values. This fact
highlights the strong sensitivity of Eckart k(T) to the barrier
height and the imaginary frequency, an issue to be discussed later.
Although the discrepancy between the Eckart and SCT
methods is small, the fundamental cause for the apparent agreement is clear. The Eckart method inherently underestimates the
tunneling probability by neglecting corner-cutting. In most cases,
this underestimation is balanced by the fact that the Eckart
function yields a potential energy curve narrower than the actual
PES, as illustrated in Figure 4, thus leading to an overestimation
of the tunneling probability. Hence, the good agreement between
Eckart and SCT transmission coecients is in many ways
fortuitous because of error cancellation in the Eckart approximation. For the same reason, the ability of the Eckart method in
reproducing the SCT results should not be generalized to other
reaction systems. For hydrogen shift reactions studied here,
however, the Eckart method is accurate.
Zhang and Dibble60 studied the impact of tunneling on
hydrogen-transfer isomerizations in n-propylperoxy radical
using multidimensional SCT calculations as well as the onedimensional, Eckart, and Wigner transmission coecients. They
reached a similar conclusion for their system. The Eckart method
works well as compared to the SCT result, but they cautioned
that the agreement may not be generalized to other systems.
Multidimensional tunneling calculations allow for the calculation of representative tunneling energies. This energy represents
the path with the greatest tunneling probability at a given
temperature. SCT representative tunneling energies for hydrogen transfers in n-heptyl radicals system are presented in Table 4.
Below 800 K, the representative tunneling energies are well
below the barrier top, and the tunneling paths can deviate
dramatically from the reaction path. As an example, Figure 5
presents the PES for 1-heptyl / 4-heptyl and the representative
tunneling energy at 300 K. Above 800 K, tunneling and reaction
paths are converging, and a one-dimensional approximation is
generally adequate. For large critical ring structures, for example,
6ps and 7ps, and at low temperatures, the representative tunneling energy is closer to the energy of the saddle point than for the
5ps or 5ss hydrogen shift. Hence, corner-cutting is enhanced for
the more strained transition state structures. It is for this reason
that SCT calculations show that the transmission coecient
increases with a decrease in the critical ring size. For example,
k(T) values are predicted to be 268 and 270 for 5ps and 5ss
H-atom shifts at 300 K, respectively. For 6ps and 7ps transitions,
k(T) values are notably smaller and equal to 30.8 and 14.3,
respectively. The dierence remains signicant until 800 K, the
temperature above which the transmission coecients are close
to unity.
We compare our SCT results to those of Zheng and Truhlar32
for 1-pentyl 5ps and 1-hexyl 5ps and 6ps H-atom shifts and those
for 1-butyl 5pp and 1-pentyl 6pp of Ratkiewitcz et al.29 As shown
in Table 5, k(300 K) values calculated for H-shift in the vemembered ring transition structure is in reasonably good agreement with literature values. For six-membered ring structures,
the k(300 K) value of Zheng and Truhlar is larger than ours by
about a factor of 3 and that of Ratkiewitcz et al. by a factor of 4.
Note that the Hessians were computed for all of the points along
the MEP in both Ratkiewitcz et al. and our SCT/VTST calculations. In particular, Ratkiewitcz et al. performed a considerable
number of force constant calculations along the MEP to ensure
convergence of SCT calculations (150 points on each side of

Table 2. CBS-QB3 Critical Energies (kcal/mol, 0 K) Computed for ips and iss Hydrogen Shift
species

7ps

6ps

5ps

n-hexyl
n-heptyl

14.7

15.3
15.0

22.1
22.0

n-octyl

14.5

15.0

22.0

n-pentyl

6ss

5ss

22.3
22.8
15.7

22.6

Table 3. Transmission Coecients Computed for 5ps


H-Atom Shift in 1-Heptyl Radical Using Wigner, Skodje and
Truhlar (S&T), Eckart, and SCT Approximations
transmission coecient, k(T)
T (K)

Wigner

S&T

Eckart

SCT

300

4.2

1.1  105

454

268

400

2.8

55.5

13.0

15.6

500

2.2

5.4

4.2

4.9

600

1.8

2.7

2.6

2.8

800

1.5

1.7

1.7

1.7

1000
1500

1.3
1.1

1.4
1.1

1.4
1.2

1.4
1.2

two methods, but we expect the present CBS-QB3 results to be


more reliable. It has also been reported that CBS-QB3 is able to
accurately predict thermochemical and kinetic data for hydrocarbon combustion.5458 In addition, Wang et al.59 studied the
thermal decomposition and isomerization of 1-hexyl radical and
showed that the CBS-QB3 calculations were able to reproduce
experimental kinetic data.
A comparison of the critical energies computed for H-shift in
dierent n-alkyl radicals is presented in Table 2. It can be seen
that for a given type of reaction, all critical energies lie within
0.3 kcal/mol of each other, indicating that the reaction energetics
is primarily a function of the ring strain energy in the critical
geometry, as will be discussed later.
Transmission Coefficients. We shall start here using the
heptyl radical as the focal point of discussion. As shown in
Table 3, transmission coefficients are calculated for 1-heptyl /
4-heptyl hydrogen transfer considering four approximations. For
T g 800 K, the transmission coefficients are close to each other
for all approximations considered. Below 800 K, however,
discrepancies become progressively larger. For 300 e T e 500 K,
transmission coefficients calculated with the SkodjeTruhlar
(S&T) approximation give the highest values, while Wigners
method produces the lowest values; both are substantially different from the higher-order SCT results (Figure 3). In comparison, the Eckart method provides a reasonably good approximation of the SCT results above 300 K. The largest discrepancy
is for the 7ps isomerization (1- to 2-heptyl) with the Eckart
k value 2.8 times that of the SCT value at 300 K. This difference is
generally smaller than that resulting from the various uncertainties in the electronic structure calculation itself. At 400 K, the
deviation becomes substantially smaller and is only about 20%.
Above 500 K, SCT and Eckart transmission coefficients are
essentially identical. It is noteworthy to mention that Ratkiewicz
et al.29 compared Eckart and SCT transmission coefficient for ipp
isomerizations, with i = 4, 5, 6, and 7. They found somewhat
larger differences between the Eckart and SCT k(T) values at
324

dx.doi.org/10.1021/jp209360u |J. Phys. Chem. A 2012, 116, 319332

The Journal of Physical Chemistry A

ARTICLE

Figure 3. Comparison of the transmission coecient k(T) for 1- and 2-heptyl radicals calculated using the Wigner, Skodje and Truhlar (S&T), Eckart,
and SCT approximations. Symbols are computed values. Lines are drawn to guide the eye.

Table 4. Representative Tunneling Energies (Ert) as a Function of Temperature Relative to the Vibrational Ground-State
Energy of the Saddle Point (E0Kq) for Hydrogen Shift in
n-Heptyl
E0Kq  Ert (kcal/mol)
T (K)

5ps

6ps

7ps

5ss

300
400

8.1
5.1

2.3
1.3

2.3
1.7

4.0
2.3

500

1.8

0.7

1.5

2.0

600

1.0

0.4

1.0

1.7

800

0.3

<0.01

0.4

1.0

1000

<0.01

<0.01

<0.01

0.4

1500

<0.01

<0.01

<0.01

<0.01

As was already mentioned above, large curvature calculations


were also performed in a somewhat qualitative manner to examine
whether the small curvature approximation is adequate for H-shift
in n-heptyl radicals. The LCT approximation describes tunneling
paths that are near and far from the minimum energy path, and
allows for a large degree of corner-cutting. Calculations show that
an LCT tunneling mechanism can be important over a small, selected
range of energies along the reaction path, but for reactions
examined here, the SCT approximation alone is suciently

Figure 4. Comparison of the potential energy () of 14 H shift in


1-heptyl and an Eckart function t (- - -).

MEP, leading to a total of 300 points due to symmetry of the


potential function for ipp reactions, with i = 5, 6, and 7).
Regardless, ve-membered ring critical geometries are shown
to have a greater tunneling tendency than the six-membered ring
transition in all of the studies.
325

dx.doi.org/10.1021/jp209360u |J. Phys. Chem. A 2012, 116, 319332

The Journal of Physical Chemistry A

ARTICLE

accurate. LCT tunneling, even for a few energy levels, does lead
to larger transmission coecients. For example, for 7ps H-atom
shift in 1-heptyl, the transmission coecient is increased by 17%
at 300 K and 41% at 250 K.
Eckart transmission coecients are known to be sensitive to
the energy barrier height and the imaginary frequency at the

saddle point. Figure 6a presents the results of sensitivity analyses


with respect to the imaginary frequency and barrier height using
1-heptyl / 4-heptyl as an example. It is seen that within (2 kcal/mol
variation, k exhibits a small sensitivity with respect to the barrier
height. At 300 K, an increase or decrease of 2 kcal/mol in the
barrier height leads to 30% change in k. What impacts the
transmission coecient the most is the uncertainty in the
imaginary frequency (i) of the critical geometry, as shown in
Figure 6b. At 600 K, k calculated with 1.1 i is 20% larger than
the reference value. At 300 K, however, deviations are as large as a
factor of 4 due to a 10% change in i.
High-Pressure Limit Rate Constants. Figure 7 presents highpressure limit rate constants computed for heptyl isomerization
using the classical TST-HR approximation with Eckart tunneling.
Molecular parameters were taken from electronic structure
calculations at the CBS-QB3 level of theory. As observed by
Zheng and Truhlar32 and Ratkiewicz et al.,29 almost no difference
was found between TST and CVT calculations, indicating
recrossing to be infrequent (a maximum deviation of 0.3%).
Figure 1 presents the computed k for reactions R1 and R2 along
with literature data. The agreement is generally good, especially
for reaction R1 considering the fairly significant scatter in the
data. The high-pressure limit rate constants of reaction R2
reported by Imbert and Marshall23 and Watkins and Ostreko18
are probably outside of the uncertainty bound of the current
calculation and those of other studies.
The high-pressure limit rate constants compare well with the
experimentally derived data of Tsang et al.,4,6,7 as illustrated in
both Figures 1 and 7. In general, the dierence is within a factor
of 2 of each other. A very interesting point is that our calculations
conrm the conclusion of Tsang et al.4 regarding the dierence in
6ps and 7ps hydrogen transfers. They proposed that the rate of
the 6ps transition is about a factor of 2 greater than the 7ps
transition and concluded that larger-ring transition structures will
possibly make a lesser contribution to isomerization because of
entropic constraints. The present results support this view. From
the PES displayed in Figure 2, it can be seen that the critical
energies for 6ps and 7ps isomerizations are similar, yet the
computed k(6ps)/k(7ps) is 5 over the temperature range
of 5002000 K.

Figure 5. Vibrationally adiabatic ground-state potential energy as a


function of reaction coordinate for the 14 hydrogen shift in 1-heptyl.
The dotted line denotes the representative tunneling energy at 300 K.

Table 5. Comparison of SCT Transmission Coecients at


300 K
reaction

reference

5ps (1-pentyl / 2-pentyl)


5ps (1-hexyl / 3-hexyl)

231
245

Zheng and Truhlar32


Zheng and Truhlar32

5pp (1-butyl / 1-butyl)

364

Ratkiewitcz et al.29

5ps (1-heptyl / 4-heptyl)

268

this work

6ps (1-hexyl / 2-hexyl)

114

Zheng and Truhlar32

6pp (1-pentyl / 1-pentyl)

38.5

Ratkiewitcz et al.29

6ps (1-heptyl / 3-heptyl)

30.1

this work

Figure 6. Sensitivity of Eckart transmission coecient computed with respect to (a) critical energy (E0Kq) and (b) the imaginary frequency (i) for 14
H-shift in 1-heptyl with the base case E0Kq and vi computed at the CBS-QB3 level of theory. Symbols are computed values. Lines are drawn to guide
the eye.
326

dx.doi.org/10.1021/jp209360u |J. Phys. Chem. A 2012, 116, 319332

The Journal of Physical Chemistry A

ARTICLE

Figure 7. Arrhenius plots of the high-pressure limit rate constant for hydrogen shifts in n-heptyl radicals.

smaller energy barrier (14.1 versus 21.3 kcal/mol). Above 1000 K,


the 7ps channel becomes less important than 5ps because of
increasing entropic resistance.
As mentioned earlier, high-pressure limit rate constants were
also computed for n-octyl radicals, again at the CBS-QB3 level of
theory using the TST-HR approximation with Eckart transmission
coecients. These computations constitute the rst systematic,
theoretical study on n-octyl isomerization. Arrhenius plots and
branching ratios for the unimolecular reactions of 1-octyl are
presented in Figure 9. For T < 800 K, rates calculated here are
within a factor of 2 from those of Tsang et al.,4 except for the 5ss
isomerization. Toward higher temperatures, discrepancies become
somewhat pronounced. At 1000 K, the rates dier by a factor of 2
for 5ps, but for 7ps the dierence is larger (a factor of 7). Qualitatively, the variations of the branching ratio as a function of
temperature follow the same trends as predicted for 1-heptyl. The
6ps H-shift is the most favored, and the 7ps H-shift is favored over
5ps below 1000 K. Again, 7ps is suppressed above 1000 K due to
entropy eects. It is worthy noting that 8ps H-shift is unimportant,
even though its critical energy (18.8 kcal/mol) is well below that of
5ps (22 kcal/mol). The isomerization is severely limited by the
entropic eect, as concluded in the earlier study.4
Modied Arrhenius parameters are presented in Table 6
for k of all H-shift reactions, from n-pentyl to n-octyl. The rate
expressions were tted in two separate temperature ranges,
because of the peculiar temperature dependence due to tunneling. The maximum absolute deviation of the ts is 6% in the low
temperature range and 3% in for high temperatures.
Reaction Rate Rules. The structurereactivity-based method2
for the estimation of rate constants of H-shift reactions has been
used for decades.61,62 Within this framework, the activation energy

Figure 8. Branching ratios computed for the thermal decomposition of


1-heptyl radical.

Branching ratios of the four unimolecular channels of 1-heptyl


radical (5ps, 6ps, and 7ps isomerizations and CC bond -scission)
are presented in Figure 8, based on the rate constants computed
at the high-pressure limit. For CC -scission, the rate constant is calculated using the critical geometry of Sirjean et al.
determined at the CBS-QB3 level of theory.55 For T < 1000 K,
the 6ps hydrogen transfer is the dominant isomerization channel.
About 1000 K, CC -scission leading to ethylene and 1-pentyl
dominates. 7ps and 5ps isomerizations are relatively minor channels, although the 7ps transition is faster below 1000 K due to a
327

dx.doi.org/10.1021/jp209360u |J. Phys. Chem. A 2012, 116, 319332

The Journal of Physical Chemistry A

ARTICLE

Figure 9. Arrhenius plots for the high-pressure limit rate constants of hydrogen shifts in n-octyl radicals. The bottom right plot represents branching
ratios computed for the thermal decomposition of 1-octyl radical.

is estimated by adding the energy required to abstract an hydrogen


atom of a given type (primary, secondary, or tertiary) to the ring
strain energy in the cyclic transition structure. The pre-exponential
(A) factor can be calculated by estimating the entropy of activation
by considering the loss or gain of internal rotor(s) and of optical
isomers. This semiempirical method relies on the group additivity
principle, with values of ring strain energies taken from those
tabulated for cycloalkanes and entropies of activation estimated
from analogous cyclic and acyclic alkanes. For example, in an early
study,63 a 4 cal/(mol K) change in the entropy was proposed for
each rotor locked into the cyclic transition structure. They derived
this value on the basis of entropy difference between n-butane
and cyclobutane (12 cal/(mol K)) and the three internal rotors
in n-butane.63 Unfortunately, the corrections were derived from
experimental data of hydrogen-transfer reactions that are inherently

subject to the tunneling effect. Additionally, the pre-exponential


factors and activation energies derived from these methods also
include the effects of tunneling. Here, we re-examine these rate
rules by decoupling the tunneling effect from the apparent
activation energies and A factors.
We assume that for a given reaction class, transmission coecients follow the same temperature dependency. This assumption is supported by the current theoretical results shown
Figure 10. As seen, the transmission coecient is primarily a
function of the ring size in the transition state and not a function
of the reactant size. Within each reaction class, the maximum
deviation is smaller than 2% for T > 500 K. Below this
temperature, the maximum deviation is 4% for all reactions
except for ve-member ring structures where a 14% maximum
deviation is observed. Tunneling can therefore be taken into
328

dx.doi.org/10.1021/jp209360u |J. Phys. Chem. A 2012, 116, 319332

The Journal of Physical Chemistry A

ARTICLE

Table 6. High-Pressure Limit Rate Expressionsa for H-Shift


in n-Alkyl Radicals Computed Using CBS-QB3 and TST/HR/
Eckart Methods
k = AT neB/RT
reaction
1-pentyl / 2-pentyl (5ps)
1-hexyl / 2-hexyl (6ps)

n
55

T (K)

1.36  10

21.02

787

300500

2.91  105

1.70

17 812

5002000

8.25  1024
2.53  105

10.61

2516

300500

1.45

11 209

5002000

504
17 898

300500
5002000

1-hexyl / 3-hexyl (5ps)

2.05  1054
6.87  105

20.69
1.59

1-heptyl / 2-heptyl (7ps)

2.25  1020

9.24

3235

300500

5.52  104

1.39

10 556

5002000

1-heptyl / 3-heptyl (6ps)

3.70  1024
4.05  105

1-heptyl / 4-heptyl (5ps)

1.23  1055
2.73  105

10.69

3235

300500

1.37

11 109

5002000

21.00

974

300500

1.65

17 718

5002000

303
18 575

300500
5002000

2-heptyl / 3-heptyl (5ss)

1.92  1054
1.24  105

20.51
1.66

1-octyl / 2-octyl (8ps)

8.65  1030

1-octyl / 3-octyl (7ps)

12.07

4658

300500

1.11  104

1.39

14 475

5002000

2.47  1021

9.49

2817

300500
5002000

6.94  10

1-octyl / 4-octyl (6ps)

3.23  1023
1.74  105

1.30

10 542

10.28

2608

300500

1.38

11 145

5002000

1-octyl / 4-octyl (5ps)

3.43  1055
2.37  105

20.84
1.65

799
17 731

300500
5002000

2-octyl / 3-octyl (6ss)

7.78  1025

10.68

2875

300500

4.38  10

2-octyl / 4-octyl (5ss)

1.49  1052
5.34  104

Figure 10. Comparison of the transmission coecients computed for


hydrogen shift in alkyl radicals using the Eckart method. O: 5ps shift in
1-pentyl, 1-hexyl, 1-heptyl, and 1-octyl and 5ss in 2-heptyl and 2-octyl.
0: 6ps shift in 1-hexyl, 1-heptyl, and 1-octyl and 6ss in 2-octyl. ]: 7ps
shift in 1-heptyl and 1-octyl.

Table 7. SCT Transmission Coecients as a Function of the


Transition State Ring Size, Fitted to a Modied Arrhenius
Expression, and Reaction Rate Rule (500 < T < 1500 K) for
n-Alkyl H-Shift Reactionsa
k(T) = CTneB/T
C

ring size

7

RSE (kcal/mol)

1.46

11 807

5002000

19.81

897

300500

5
6c

9.28  10
1.51  104

1.704
1.068

2431
1711

10.4
3.4

1.69

18 491

5002000

7d

1.12  104

1.118

1600

2.9

Units are cm , s, cal, mol.


Eabs (kcal/mol)

account within the framework of a semiempirical correlation by


using a modied Arrhenius equation leading to the parameter
presented in Table 7. Here, small curvature transmission coecients determined for the n-heptyl system were used as reference
data. The temperature range for the t was chosen to minimize
the errors induced by the tting procedure. In all cases, the tting
error is <5%.
Hydrogen transfer reactions are fundamentally internal H-atom
abstraction reaction. As we discussed earlier, the activation
energies can be estimated using the following equation:
E Eabs RSE

E = Eabs + RSE

ps

ss

11.7

12.6

(Sq) (cal/(mol K))

ekB T S n
h
R

ps

ss

3.61

4.56

A reaction path degeneracy of 2 is included in the rate rules. b 500 e T e


1500 K. c 400 e T e 1500 K. d 300 e T e 1500 K. e n is the number of
internal rotors lost between the cyclic transition state structure and the reactant.
a

19

where Eabs is the critical energy for the abstracting a secondary H


atom by a primary (in the case of ips processes) or a secondary
(iss) radical, and RSE is the ring strain energy. Using a mean
value for the critical energy of a given reaction class (Table 2) and
critical energies for the H-abstraction of a secondary H-atom by
p-C3H7 and i-C3H7 radicals from propane, RSE values are
estimated, as shown in Table 8. Here, the critical energy of the
abstraction reactions was computed at the CBS-QB3 level of
theory. As seen, RSE values derived for the transition state
structure are higher than the corresponding ring strain energies
of cycloalkanes (also listed in Table 8) except for the sevenmembered ring structure. For ve- and six-membered rings, our

Table 8. Ring Strain Energies (kcal/mol, 0 K) Determined at


the CBS-QB3 Level of Theory (See Text and Eq 20)

329

RSE in cycloalkanes54

reaction class

Eabs

RSE

5ps

22.1

11.7a

10.4

7.1

6ps

15.1

11.7a

3.4

0.7

7ps

14.6

11.7a

2.9

6.8

5ss
6ss

22.7
15.7

12.6b
12.6b

10.1
3.1

7.1
0.7

H-abstraction of C3H8 by p-C3H7. b H-abstraction of C3H8 by i-C3H7.


dx.doi.org/10.1021/jp209360u |J. Phys. Chem. A 2012, 116, 319332

The Journal of Physical Chemistry A

ARTICLE

Table 9. Contributions of Internal Rotor Losses to the Entropy of Activation in iab Hydrogen Shift
per rotor entropy loss ((Sq) = Sq/n,

number of internal
reaction channel

rotors lost (n)

S298Kq (cal/(mol K))

cal/(mol K))

1-pentyl / 2-pentyl (5ps)

10.00

1-hexyl / 2-hexyl (6ps)

14.02

3.33
3.51

1-hexyl / 3-hexyl (5ps)

10.23

3.41

1-heptyl / 2-heptyl (7ps)

17.79

3.56

1-heptyl / 3-heptyl (6ps)

14.57

3.64

1-heptyl / 4-heptyl (5ps)

11.27

3.76

2-heptyl / 3-heptyl (5ss)


1-octyl / 2-octyl (8ps)

3
6

13.08
19.83

4.36
3.31

1-octyl / 3-octyl (7ps)

18.80

3.76

1-octyl / 4-octyl (6ps)

16.08

4.02

1-octyl / 4-octyl (5ps)

11.57

3.86

2-octyl / 3-octyl (6ss)

18.07

4.52

2-octyl / 4-octyl (5ss)

14.41

4.80

mean for ips

3.61 ( 0.23

mean for iss

4.56 ( 0.22

For most cases, the deviation is well within the estimated


uncertainty of the theoretical calculation (around a factor of 2).

4. CONCLUSION
The role of quantum tunneling in hydrogen-transfer isomerizations of linear alkyl radicals was studied in detail. Transmission
coecients are shown to require a multidimensional treatment
below 800 K. Only above 800 K is a one-dimensional treatment
appropriate. The Eckart method is shown to reproduce the
multidimensional transmission coecients over the entire temperature range, but the agreement is due to a favorable error
cancellation. The inability of the Eckart approach to account for
higher-dimensional tunneling eect is compensated by the
rigidity of the Eckart function, leading to a potential energy
curve narrower than the actual PES. Calculations show that
below 600 K the Eckart transmission coecient is highly
sensitive to the value of the imaginary frequency, and hence is
subject to the uncertainty in the electronic structure calculation.
Regardless, high-pressure limit rate constants calculated using
the classical transition state theory with treatment of internal
rotation and the use of Eckart transmission coecients and a PES
determined at the CBS-QB3 level of theory are in good agreement with literature data.
The present results are consistent with Tsang et al., who
concluded that a large body of literature data at low temperatures
(with small A factor and activation energy) can be reconciled
with high-temperature data by taking into account quantum
tunneling. Our calculations also conrm their observation that
tunneling is more pronounced with ve-membered ring transition structures than with six-membered ring structures. In
addition, hydrogen transfer through an eight-membered ring
transition structure is not competitive due to prohibitive entropy
eects.
The systematic study of isomerization reactions from n-pentyl
to n-octyl led to a large set of theoretical kinetic data that can be
rationalized within a reaction class approach. It is shown that the
transmission coecient can be treated as an explicit structure/
reactivity correlation parameter. This approach allows for a fast
and suciently accurate estimation of the impact of tunneling on

Figure 11. Comparison of the high-pressure limit rate constants ()


and rate-rule estimates ( 3 3 3 ) for hydrogen shift in n-heptyl radials.

values are consistently 2.5 kcal mol1 above those of cyclic


alkanes. Hence, using the ring strain energy of cycloheptane
would lead to an overestimation of the activation energy.
Contributions of the loss of internal rotation to the activation
entropy for each hydrogen-transfer reaction studied in this work
are presented in Table 9. Only values at 298 K are reported here.
Temperature was found to have a negligible eect on the entropy
of activation. For ips hydrogen transfers, the loss of one internal
rotor causes the entropy of activation to decrease by 3.61 ( 0.23
cal/(mol K), while a mean value of 4.56 ( 0.23 cal/(mol K)
per rotor was calculated for iss hydrogen shift. The dierence is
not surprising considering that ips involves CH2CH2 and
CH2CH2 3 rotors, but iss involves CH2CH2 internal
rotations only.
The recommended hydrogen-transfer rate rule is summarized
in Table 7. Comparisons between the theoretical k values and
the rate-rule estimations are shown in Figure 11 for hydrogen
shift in n-heptyl radicals. For the temperature range considered
(5001500 K), the maximum deviation is about a factor of 2.5.
330

dx.doi.org/10.1021/jp209360u |J. Phys. Chem. A 2012, 116, 319332

The Journal of Physical Chemistry A

ARTICLE

the rate constant. It can be used also to extrapolate kinetic data


obtained in a limited temperature range. A rate rule was proposed
for hydrogen shift in linear alkyl radicals. In this rule, transmission coecients are calculated using multidimensional tunneling
for a representative reaction class. The rate constant without
tunneling may be estimated by considering the loss or gain
of internal rotations in the transition state structure for preexponential factors. The apparent activation energy may be estimated from that of typical H-abstraction by an alkyl radical
corrected for ring strain. The rate rule proposed here is shown to
reproduce the theoretical high-pressure limit rate constant to within
a factor of 2 for almost the entire range of temperatures considered.

(9) Wang, H.; Frenklach, M. Combust. Flame 1997, 110, 173221.


(10) Wang, H. Proc. Combust. Inst. 2011, 33, 4167.
(11) Frenklach, M.; Wang, H. Symp. (Int.) Combust. 1991, 23, 1559
1566.
(12) You, X.; Egolfopoulos, F. N.; Wang, H. Proc. Combust. Inst.
2009, 32, 403410.
(13) McMahon, R. J. Science 2003, 299, 833834.
(14) Frey, H. M.; Walsh, R. Chem. Rev. 1969, 69, 103124.
(15) Endrenyi, L.; Le Roy, D. J. J. Phys. Chem. 1966, 70, 40814084.
(16) Watkins, K. W. J. Am. Chem. Soc. 1971, 93, 63556359.
(17) Watkins, K. W. Can. J. Chem. 1971, 50, 37383740.
(18) Watkins, K. W.; Ostreko, L. A. J. Phys. Chem. 1969, 73, 2080
2083.
(19) Hardwidge, E. A.; Larson, C. W.; Rabinovitch, B. S. J. Am. Chem.
Soc. 1970, 92, 32783283.
(20) Mintz, K. J.; Leroy, D. J. Can. J. Chem. 1973, 51, 35343538.
(21) Dobe, S.; Berces, T.; Reti, F.; Marta, F. Int. J. Chem. Kinet. 1987,
19, 895921.
(22) Marshall, R. M. Int. J. Chem. Kinet. 1990, 22, 935950.
(23) Imbert, F. E.; Marshall, R. M. Int. J. Chem. Kinet. 1987,
19, 81103.
(24) Yamauchi, N.; Miyoshi, A.; Kosaka, K.; Koshi, M.; Matsui, N.
J. Phys. Chem. A 1999, 103, 27232733.
(25) Miyoshi, A.; Widjaja, J.; Yamauchi, N.; Koshi, M.; Matsui, H.
Proc. Combust. Inst. 2002, 29, 12851293.
(26) Matheu, D. M.; Green, W. H.; Grenda, J. M. Int. J. Chem. Kinet.
2003, 35, 95119.
(27) Hayes, C. J.; Burgess, D. R., Jr. J. Phys. Chem. A 2009, 113,
24732482.
(28) Bankiewicz, B.; Huynh, L. K.; Ratkiewicz, A.; Truong, T. N.
J. Phys. Chem. A 2009, 113, 15641573.
(29) Ratkiewicz, A.; Bankiewicz, B.; Truong, T. N. Phys. Chem.
Chem. Phys. 2010, 12, 1098810995.
(30) Jitariu, L. C.; Jones, L. D.; Robertson, S. H.; Pilling, M. J.;
Hillier, I. H. J. Phys. Chem. A 2003, 107, 86078617.
(31) Truhlar, D. G.; Issacson, A. D.; Garrett, B. C. In Theory of
Chemical Dynamics; Baer, M., Ed.; CRC Press: Boca Raton, FL, 1985; pp
68135.
(32) Zheng, J.; Truhlar, D. G. J. Phys. Chem. A 2009, 113, 11919
11925.
(33) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.;
Robb, M. A.; Cheeseman, J. R.; Montgomery, J. A.; Vreven, T.; Kudin,
K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.;
Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.;
Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa,
J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.;
Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.; Adamo, C.;
Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.;
Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.;
Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich,
S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.;
Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.;
Cliord, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.;
Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Laham, A.;
Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson,
B.; Chen, W.; Wong, M. W.; Gonzalez, C.; Pople, J. A. Gaussian 03,
revision B.03; Gaussian, Inc.: Pittsburgh, PA, 2004.
(34) Shao, Y.; Molnar, L. F.; Jung, Y.; Kussmann, J.; Ochsenfeld, C.;
Brown, S. T.; Gilbert, A. T. B.; Slipchenko, L. V.; Levchenko, S. V.;
ONeill, D. P.; DiStasio, R. A.; Lochan, R. C.; Wang, T.; Beran, G. J. O.;
Besley, N. A.; Herbert, J. M.; Lin, C. Y.; Van Voorhis, T.; Chien, S. H.;
Sodt, A.; Steele, R. P.; Rassolov, V. A.; Maslen, P. E.; Korambath, P. P.;
Adamson, R. D.; Austin, B.; Baker, J.; Byrd, E. F. C.; Dachsel, H.;
Doerksen, R. J.; Dreuw, A.; Dunietz, B. D.; Dutoi, A. D.; Furlani, T. R.;
Gwaltney, S. R.; Heyden, A.; Hirata, S.; Hsu, C. P.; Kedziora, G.;
Khalliulin, R. Z.; Klunzinger, P.; Lee, A. M.; Lee, M. S.; Liang, W.; Lotan,
I.; Nair, N.; Peters, B.; Proynov, E. I.; Pieniazek, P. A.; Rhee, Y. M.;
Ritchie, J.; Rosta, E.; Sherrill, C. D.; Simmonett, A. C.; Subotnik, J. E.;

ASSOCIATED CONTENT

bS

Supporting Information. Optimized geometries at the


B3LYP/6-311G(2d,d,p) level of theory for all species. Hindrance
potentials and barrier heights at the CBS-QB3 level of theory.
Forward and reverse reaction barrier heights (E1 and E1) at 0 K
along with imaginary frequencies (i) of the transition states and
calculated characteristic lengths. SCT transmission coecients
from 200 to 2000 K for all n-heptyl H-atom shifts. This material is
available free of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION
Corresponding Author

*Phone: (33) 383175202 (B.S.), (213) 740-0499 (H.W.). E-mail:


baptiste.sirjean@ensic.inpl-nancy.fr (B.S.), haiw@usc.edu (H.W.).
Present Addresses

Laboratoire Reactions et Genie des Procedes, Nancy Universite,


CNRS, BP 20451, 1 rue Grandville, 54001 Nancy, France.

ACKNOWLEDGMENT
This work was supported by the U.S. Air Force Oce of
Scientic Research (AFOSR Grant Nos. FA9550-07-1-0168 and
FA9550-08-1-0040) and by the Combustion Energy Frontier
Research Center (CEFRC), an Energy Frontier Research Center
funded by the U.S. Department of Energy, Oce of Science, Oce
of Basic Energy Sciences under Award No. DE-SC0001198. Part
of this work was performed using HPC resources from GENCICINES (Grant 2011086686).
REFERENCES
(1) Viskolcz, B.; Lendvay, G.; Kortvelyesi, T.; Seres, L. J. Am. Chem.
Soc. 1996, 118, 30063009.
(2) Benson, S. W. Thermochemical Kinetics, 2nd ed.; Wiley: New York,
1976.
(3) Tsang, W. J. Phys. Chem. A 2006, 110, 85018509.
(4) Tsang, W.; Awan, I.; McGovern, S.; Manion, J. A. Soot Precursors from Real Fuels: The Unimolecular Reactions of Fuel Radicals.
In Combustion Generated Fine Carbon Particles; Bockhorn, H., DAnna,
A., Sarom, A. F., Wang, H., Eds.; KIT Scientic Publishing: Karlsruhe,
2009; pp 5574.
(5) Tsang, W.; McGivern, W. S.; Manion, J. A. Proc. Combust. Inst.
2009, 32, 131138.
(6) Tsang, W.; Walker, J. A.; Manion, J. A. Symp. (Int.) Combust.
1998, 1, 135142.
(7) Tsang, W.; Walker, J. A.; Manion, J. A. Proc. Combust. Inst. 2007,
31, 141148.
(8) Richter, H.; Howard, J. B. Prog. Energy Combust. Sci. 2000,
26, 565608.
331

dx.doi.org/10.1021/jp209360u |J. Phys. Chem. A 2012, 116, 319332

The Journal of Physical Chemistry A

ARTICLE

Woodcock, H. L.; Zhang, W.; Bell, A. T.; Chakraborty, A. K.; Chipman,


D. M.; Keil, F. J.; Warshel, A.; Hehre, W. J.; Schaefer, H. F.; Kong, J.;
Krylov, A. I.; Gill, P. M. W.; Head-Gordon, M. Phys. Chem. Chem. Phys.
2006, 8, 31723191.
(35) Becke, A. D. J. Chem. Phys. 1993, 98, 56485652.
(36) Lee, C. T.; Yang, W. T.; Parr, R. G. Phys. Rev. B 1988, 37,
785789.
(37) Montgomery, J. A.; Frisch, M. J.; Ochterski, J. W.; Petersson,
G. A. J. Chem. Phys. 1999, 110, 28222827.
(38) Vansteenkiste, P.; Van Speybroeck, V.; Marin, G. B.; Waroquier,
M. J. Phys. Chem. A 2003, 107, 31393145.
(39) Pitzer, K. S.; Gwinn, W. D. J. Chem. Phys. 1942, 10, 428440.
(40) Mokrushin, V.; Tsang, W. Chemrate v.1.5.2; National Institute
of Standards and Technology: Gaithersburg, MD, 2006.
(41) Pitzer, K. S. J. Chem. Phys. 1946, 14, 239243.
(42) Miller, W. H. J. Am. Chem. Soc. 1979, 101, 68106814.
(43) Wigner, E. Z. Phys. Chem. 1932, 19, 203.
(44) Skodje, R. T.; Truhlar, D. G. J. Phys. Chem. 1981, 85, 624628.
(45) Eckart, C. Phys. Rev. 1930, 35, 13031309.
(46) Lu, D. h.; Truong, T. N.; Melissas, V. S.; Lynch, G. C.; Liu, Y. P.;
Garrett, B. C.; Steckler, R.; Isaacson, A. D.; Rai, S. N.; Hancock, G. C.;
Lauderdale, J. G.; Joseph, T.; Truhlar, D. G. Comput. Phys. Commun.
1992, 71, 235262.
(47) Liu, Y. P.; Lynch, G. C.; Truong, T. N.; Lu, D. H.; Truhlar,
D. G.; Garrett, B. C. J. Am. Chem. Soc. 1993, 115, 24082415.
(48) Fernandez-Ramos, A.; Truhlar, D. G. J. Chem. Phys. 2001,
114, 14911496.
(49) Corchado, J. C.; Chuang, Y.-Y.; Fast, P. L.; Hu, W.-P.; Liu, Y. P.;
Lynch, G. C.; Nguyen, K. A.; Jackels, C. F.; Fernandez-Ramos, A.;
Ellingson, B. A.; Lynch, B. J.; Zheng, J.; Melissas, V. S.; Villa, J.; Rossi, I.;
Coiti~no, E. L.; Pu, J.; Albu, T. V.; Steckler, R.; Garrett, B. C.; Isaacson,
A. D.; Truhlar, D. G. POLYRATEversion 9.7; University of Minnesota:
Minneapolis, MN, 2007.
(50) Corchado, J. C.; Chuang, Y.-Y.; Coiti~no, E. L.; Ellingson, B. A.;
Zheng, J.; Truhlar, D. G. GAUSSRATEversion 9.7; University of
Minnesota: Minneapolis, MN, 2007.
(51) Johnston, H. S.; Heicklen, J. J. Phys. Chem. 1962, 66, 532533.
(52) Sirjean, B.; Dames, E.; Sheen, D. A.; You, X.-Q.; Sung, C.;
Holley, A. T.; Egolfopoulos, F. N.; Wang, H.; Vasu, S. S.; Davidson,
D. F.; Hanson, R. K.; Pitsch, H.; Bowman, C. T.; Kelley, A.; Law, C. K.;
Tsang, W.; Cernansky, N. P.; Miller, D. L.; Violi, A.; Lindstedt, R. P. A
high-temperature chemical kinetic model of n-alkane oxidation. JetSurF
version 1.0, 2009.
(53) Holley, A. T.; You, X. Q.; Dames, E.; Wang, H.; Egolfopoulos,
F. N. Proc. Combust. Inst. 2009, 32, 11571163.
(54) Sirjean, B.; Glaude, P. A.; Ruiz-Lopez, M. F.; Fournet, R. J. Phys.
Chem. A 2006, 110, 1269312704.
(55) Sirjean, B.; Glaude, P. A.; Ruiz-Lopez, M. F.; Fournet, R. J. Phys.
Chem. A 2008, 112, 1159811610.
(56) Sirjean, B.; Glaude, P. A.; Ruiz-Lopez, M. F.; Fournet, R. J. Phys.
Chem. A 2009, 113, 69246935.
(57) Miyoshi, A. Int. J. Chem. Kinet. 2010, 42, 273288.
(58) Asatryan, R.; Bozzelli, J. W. J. Phys. Chem. A 2010, 114,
76937708.
(59) Wang, F.; Cao, D. B.; Liu, G.; Ren, J.; Li, Y. W. Theor. Chem. Acc.
2010, 126, 8798.
(60) Zhang, F.; Dibble, T. S. Phys. Chem. Chem. Phys. 2011,
13, 1796917977.
(61) Curran, H. J.; Gauri, P.; Pitz, W. J.; Westbrook, C. K. Combust.
Flame 1998, 114, 149177.
(62) Buda, F.; Bounaceur, R.; Warth, V.; Glaude, P. A.; Fournet, R.;
Battin-Leclerc, F. Combust. Flame 2005, 142, 170186.
(63) ONeal, H. E.; Benson, S. W. J. Phys. Chem. 1967, 71, 29032921.

332

dx.doi.org/10.1021/jp209360u |J. Phys. Chem. A 2012, 116, 319332

Вам также может понравиться