View Article Online / Journal Homepage / Table of Contents for this issue
PCCP
Dynamic Article Links
Cite this: Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
PAPER
www.rsc.org/pccp
Automatic estimation of pressure-dependent rate coefficientsw
Downloaded by Fritz Haber Institut der Max Planck Gesellschaft on 21 November 2012
Published on 07 December 2011 on http://pubs.rsc.org | doi:10.1039/C1CP22765C
Joshua W. Allen,a C. Franklin Goldsmithab and William H. Green*a
Received 30th August 2011, Accepted 27th October 2011
DOI: 10.1039/c1cp22765c
A general framework is presented for accurately and efficiently estimating the phenomenological
pressure-dependent rate coefficients for reaction networks of arbitrary size and complexity using
only high-pressure-limit information. Two aspects of this framework are discussed in detail.
First, two methods of estimating the density of states of the species in the network are presented,
including a new method based on characteristic functional group frequencies. Second, three
methods of simplifying the full master equation model of the network to a single set of
phenomenological rates are discussed, including a new method based on the reservoir state and
pseudo-steady state approximations. Both sets of methods are evaluated in the context of the
chemically-activated reaction of acetyl with oxygen. All three simplifications of the master
equation are usually accurate, but each fails in certain situations, which are discussed. The new
methods usually provide good accuracy at a computational cost appropriate for automated
reaction mechanism generation.
1
Introduction
Thermal unimolecular reactions—the set of reactions involving
a single reactant molecule—are unusual in that they require an
inert third body to proceed. The third body provides or
removes the energy needed for reaction via collisions with
the reactant molecule. The rate of these collisions depends on
the concentration of the inert, which in turn is related to the
pressure of the system; therefore, under conditions where this
collision rate is rate-limiting, the observed phenomenological
rate coefficient k(T,P) is a function of both temperature T and
pressure P.
An activated species (isomer) can be formed either as the
product of an association reaction (chemical activation) or via
collisional excitation (thermal activation). Once activated,
multiple isomerization and dissociation reactions may become
competitive with one another and with collisional stabilization; these combine to form a network of unimolecular
reactions. Often the reactive events occur more rapidly than
collision events, and an excited molecule will traverse multiple
reactive events—appearing to ‘‘skip’’ over intermediate species—
before being collisionally stabilized. This suggests that there is
a net reaction rate from each isomer and product to every
other isomer and product set in the network, not just those
directly adjacent.
a
Dept. of Chemical Engineering, Massachusetts Institute of
Technology, 77 Massachusetts Ave, Cambridge, MA 02139, USA.
E-mail:
[email protected]; Fax: 617-324-0066; Tel: 617-253-4580
b
Dept. of Inorganic Chemistry, Fritz Haber Institute, Berlin, 14195,
Germany
w Electronic supplementary information (ESI) available. See DOI:
10.1039/c1cp22765c
This journal is
c
the Owner Societies 2012
Our interest in unimolecular reaction networks is colored
by our intended application: automatic reaction mechanism
generation. Our automatic mechanism generation code, RMG,1
utilizes a rate-based approach to model enlargement, and it is
therefore important to account for the pressure dependence of
chemically-activated reaction networks in order for RMG to
generate a valid mechanism. RMG often considers thousands
of these reaction networks in the course of constructing a
reaction mechanism, so we require a method that is fully
automatable, successful over a wide range of conditions,
reasonably accurate, and not too computationally intensive.
At the time it is making an estimate of k(T,P), RMG has only
limited information available about the molecules and reactions, as discussed below. This makes it more challenging to
accurately estimate all the k(T,P).
The importance of bimolecular collisions in unimolecular
reactions was first proposed by Lindemann in 1922.2 It was
soon recognized by Hinshelwood and others that a rigorous
treatment of these processes required consideration of molecular energy levels.3 The RRKM expression for the microcanonical rate coefficient k(E) was derived in the early
1950s.4–6 In the late 1950s master equation models of chemical
systems began appearing,7–11 including an early linear integraldifferential equation formulation by Widom.12 Analytical
solutions for a variety of simple models soon followed,13–15
as did the first numerical approaches.16 Numerical methods—
which are required for complex unimolecular reaction networks—
became much more attractive in the 1970s with the appearance
of new algorithms, including Gear’s method for solving stiff
systems of ordinary differential equations17 and efficient
algorithms for calculating the density of states.18–20 In the
1990s computing power had increased to the point where it
Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
1131
Downloaded by Fritz Haber Institut der Max Planck Gesellschaft on 21 November 2012
Published on 07 December 2011 on http://pubs.rsc.org | doi:10.1039/C1CP22765C
View Article Online
was practical to solve them numerically by discretizing the
integrals over energy.
Methods for directly solving the discretized master equations
include integration using a stiff ODE solver and stochastic
simulation using the Gillespie method. Both of these methods
have the benefit of not introducing additional approximations
(in the latter only if sufficient sampling of events is taken).
However, both methods suffer when wide ranges of time scales
are encountered, as they often are in chemical kinetics.
Furthermore, with either approach it is not clear how to
extract the time-independent rate constants k(T,P) from the
full solution. Barker et al. have developed the stochastic
approach into the software package MultiWell.21
Over the last ten to fifteen years a number of numerical
techniques have been developed with the goal of simplifying a
full master equation formulation into a single set of phenomenological rate coefficients k(T,P). The methods vary in the
simplifying assumptions made and the numerical procedures
used, which in turn affects the computational expense required
and the accuracy and robustness of the result. A sampling of
methods and associated software packages is provided in the
following paragraphs.
Several approaches are based on computing the eigenmodes
of the full master equation matrix. When the conventional
phenomenological-kinetics description of the chemical system
is applicable, certain ‘‘chemically-significant’’ eigenmodes
become separable from the large number of internal energy
relaxation eigenmodes. The chemically-significant eigenvalues
match the eigenvalues of the matrix of phenomenological
rates, and therefore can be used to extract values for each
k(T,P), although this may not be straightforward.22 Eigenvalue determination is computationally expensive and numerically challenging for large, stiff systems; e.g. at low temperatures
and high pressures, solvers often return nonphysical or complex
eigenvalues. There are multiple software packages that utilize
eigenvalue approaches, including UNIMOL,23 ChemRate,24
Variflex,25 and the recently-released MESMER.26
On the opposite end of the scale are methods based around
the strong collision approximation, wherein all collisions are
assumed to irreversibly stabilize the energized species. This
decoupling of the energy grains in the system is a huge
computational savings, at the cost of neglecting the effects of
collisions which cause small energy changes. Despite this gross
simplification, the modified strong collision method appears to
provide reasonably accurate results, typically within an order
of magnitude of the true rate coefficients.27 This may be
partially due to the modification of the collision frequency
by a temperature-dependent efficiency factor.28–30 For multiwell systems, this method was codified by Chang, Bozzelli, and
Dean into the software package CHEMDIS.31 The modified
strong collision method is implemented in the automatic
mechanism generation software package XMG, and in early
and current versions of RMG.
More recently, Green and Bhatti have proposed a new
approximate method for computing k(T,P) based around the
pseudo steady-state approximation at high energies coupled
with the approximation of a Boltzmann population of low
energy levels.32 This method maintains the collision model of
the full analysis for the important high-energy states, so a
1132
Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
more accurate result is expected when compared to the
modified strong collision method. However, it is not clear
yet just how much more expensive and more accurate this
method is. Also, the multiple-well formulation of the reservoir
state method as described by Green and Bhatti is cumbersome,
requiring the method to be repeated using each isomer well
and reactant set as a starting point. The present work will
develop a formulation that does not require this repetition.
A key input to all k(T,P) calculations is the density of states
for each species. In the context of automated reaction mechanism generation it is not yet possible to determine this
information by performing quantum chemistry calculations
on every species, as these are still too slow for large molecules,
and nontrivial to automate. Bozzelli and coworkers developed a
method for estimating the density of states from heat capacity
data (obtained from group additivity estimates) by fitting three
pseudofrequency-degeneracy pairs.33 The fitted frequencies do
not necessarily reflect the real vibrational degrees of freedom of
the molecule, and the harmonic-oscillator form is not a good
representation for hindered internal rotors.
The primary objective of the current work was to develop a
general framework for automatic estimation of pressuredependent rate coefficients in reaction networks of arbitrary
size and complexity. Within this framework, various methods
of estimating the density of states and extracting phenomenological rate constants from the master equation matrix could
be developed and evaluated. In the former, a new method
based on characteristic functional-group frequencies was
created to more accurately reflect the real molecular degrees
of freedom, including hindered rotors. In the latter, the
chemically-significant eigenvalues, modified strong collision,
and reservoir state methods were directly compared using
identical input parameters.
Section 2 provides a brief review of the formulation of the
master equation model. Section 3 describes our framework for
estimating the phenomenological rate coefficients for unimolecular reaction networks. Section 4 describes the methods
used to estimate the density of states, focusing on a new
method based on characteristic functional group frequencies.
Section 5 describes the methods used to estimate the phenomenological rate coefficients from the full master equation
formulation. Section 6 evaluates the methods of the previous
sections using a case study: the reaction of acetyl radical with
oxygen, a process important in atmospheric chemistry.
2
Background: the master equation
Descriptions of the full master equation formulation are quite
ubiquitous,34–38 and so only a brief summary will be given
here.
The system of interest is a set of chemically reactive molecular configurations—local minima on a potential energy
surface—divided into unimolecular isomers and bimolecular
reactants or products. In our vernacular, reactants can associate
to form an isomer, while such association is neglected for
products. These configurations are connected by chemical
reactions to form a network. The system also consists of an
excess of inert gas M, representing a thermal bath; this allows
This journal is
c
the Owner Societies 2012
Downloaded by Fritz Haber Institut der Max Planck Gesellschaft on 21 November 2012
Published on 07 December 2011 on http://pubs.rsc.org | doi:10.1039/C1CP22765C
View Article Online
for neglecting all collisions other than those between an isomer
and the bath gas.
An isomer molecule at sufficiently high internal energy E
can be transformed by a number of possible events:
The isomer molecule can collide with any other molecule,
resulting in a change in energy from E to E 0
The isomer molecule can isomerize to an adjacent isomer
at constant E
The isomer molecule can dissociate into any directly
connected bimolecular reactant or product channel
The dependent variables describing the isomers are pi(E,t),
the population distribution of isomer i over energy E at time t.
From statistical mechanics we expect a Boltzmann distribution
bi(E,T) to result at long times:
lim pi ðE; tÞ ¼ xi1
t!1
ri ðEÞebE
¼ xi1 bi ðE; TÞ
Qi ðbÞ
ð1Þ
Above, xiN is the total population of isomer i at equilibrium,
ri(E) and Qi(b) are the rovibrational density of states and
corresponding partition function of isomer i, and b (kBT)1.
Assuming that inelastic collisions are much more common
than reactive collisions, we can treat the bimolecular reactants
as thermalized, and represent the concentrations of the nth
reactant configuration by ynA(t) and ynB(t), where A and B
distinguish between the two reactants.
Neglecting the dependencies of the microcanonical rates and
other quantities on the angular momentum quantum number
J or any other quantum numbers besides E,39 and neglecting
reactions which do not proceed through one of the isomers,
the master equation can be written as
Z 1
d
pi ðE; tÞ ¼ oi ðT; PÞ
Pi ðE; E 0 ; TÞpi ðE 0 ; tÞdE 0
dt
0
oi ðT; PÞpi ðE; tÞ
þ
N
isom
X
þ
N
reac
X
jai
n¼1
kij ðEÞpj ðE; tÞ
jai
kji ðEÞpi ðE; tÞ
ynA ðtÞynB ðtÞfin ðEÞbn ðE; TÞ
Nreac þNprod
X
N
isom
X
n¼1
gni ðEÞpi ðE; tÞ
ð2Þ
d
d
ynA ðtÞ ¼
ynB ðtÞ
dt
dt
Z 1
N
isom
X
gni ðEÞpi ðE; tÞdE
¼
i¼1
0
N
isom
X
i¼1
ynA ðtÞynB ðtÞ
Z
0
1
ð3Þ
fin ðEÞbn ðE; TÞdE
where oi(T,P) is the collision frequency; P(E,E 0 ,T) is the
probability of collisional transfer from energy E 0 to E; kij(E),
fin(E), and gni(E) are the microcanonical rate coefficients
for isomerization, association, and dissociation, respectively;
This journal is
c
the Owner Societies 2012
bn(E,T) is the Boltzmann distribution for bimolecular reactant
channel n; and Nisom, Nreac, and Nprod are the numbers
of isomers, bimolecular reactant channels, and bimolecular
product channels, respectively. In eqn (2), the first pair of
terms correspond to collision, the second pair to isomerization, and the final pair to association/dissociation. Eqn (2)
applies to isomers, while eqn (3) applies to bimolecular
reactants. Note that there are almost always additional reactions creating and destroying all the species, so eqn (2) and (3)
are for an idealized situation.
There is some subtlety to the formulation of the master
equation, in that some treatments are based on the total
energy, while others use the active internal energy. This has
been discussed in the literature previously.40,41 This work
utilizes the one-dimensional total energy model as described
by Miller and Klippenstein. In particular, the densities of
states contain all of the vibrational and rotational modes of
the system, which implies that we are treating all rotational
modes as active; this approximation appears to be increasingly
accurate as temperature increases.41
Eqn (2) and (3) are nonlinear, both due to the presence of
the bimolecular reactant terms and because both oi and
P(E,E 0 ) depend on the composition, which is changing with
time. The rate coefficients can be derived from considering the
pseudo-first-order situation where ynA(t) { ynB(t), and all y(t)
are negligible compared to the bath gas M. From these
assumptions the changes in o,P(E,E 0 ), and all ynB can be
neglected, which yields a linear equation system.
To extract the phenomenological rate coefficients numerically, it is helpful to discretize the energy E into Ngrains grains
{Er}. This converts the linear integro-differential equations
into a system of first-order ordinary differential equations with
the form
d
P ¼ MP
dt
2
M1
6
6 K21
6
6
6 .
6 .
6 .
6
M¼6
6 ðg ÞT
6 11
6
6
6 ðg21 ÞT
6
4
..
.
3
2
p1
7
6
6 p 7
6 2 7
7
6
6 . 7
6 . 7
6 . 7
7
6
P¼6
7
6y 7
6 1A 7
7
6
7
6
6 y2A 7
7
6
5
4
..
.
K12
...
F11 b1 y1B
F12 b2 y2B
...
M2
...
F21 b1 y1B
F22 b2 y2B
..
.
..
.
..
.
..
.
ðg12 ÞT
...
h1
0
ðg22 ÞT
...
0
h2
..
.
..
..
.
..
.
7
...7
7
7
.. 7
7
.7
7
7
...7
7
7
7
...7
7
5
..
.
.
Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
3
ð4Þ
1133
View Article Online
where the elements of the vectors pi are such that (pi)r = pi(Er).
The diagonal matrices Kij and Fin and the vectors gni contain
the microcanonical rate coefficients for isomerization, association, and dissociation, respectively:
R Er þDEr =2
1
ðKij Þrs ¼ DEr Er DEr =2 kij ðEÞdE r ¼ s
ð5aÞ
0
ras
Downloaded by Fritz Haber Institut der Max Planck Gesellschaft on 21 November 2012
Published on 07 December 2011 on http://pubs.rsc.org | doi:10.1039/C1CP22765C
ðFin Þrs ¼
R Er þDEr =2
1
DEr Er DEr =2 fin ðEÞdE
0
ðgni Þr ¼
1
DEr
Z
Er þDEr =2
Er DEr =2
r¼s
ras
ð5bÞ
gni ðEÞdE
ð5cÞ
Above, r is the index for the discretized energies of isomer i,
and s is the index for the discretized energies of isomer j or
bimolecular channel n. The matrices Mi represent the collisional energy transfer probabilities of isomer i minus the rates
of reactive loss to other isomers and to reactants and products:
8
isom
>
NX
>
> oi ðT; PÞ ðPi Þ 1
ðKij Þrs
>
rr
>
>
>
jai
>
<
r¼s
ðMi Þrs ¼
Nreac þNprod
X
>
>
>
>
ðgni Þr
>
>
>
n¼1
>
:
oi ðT; PÞðPi Þrs
ras
ð6Þ
The collisional energy transfer probabilities (Pi)rs for isomer i
are given by
Z Er þDEr =2 Z Es þDEs =2
1
ðPi Þrs ¼
Pi ðE; E 0 ; TÞdE 0 dE
DEr DEs Er DEr =2 Es DEs =2
ð7Þ
The scalars hn are simply the total rate coefficient for loss of
reactant channel n due to chemical reactions:
hn ¼
N
grains
isom NX
X
i¼1
r¼1
ynB ðFin Þrr bn ðEr Þ
ð8Þ
3 A general framework for processing reaction
networks
The procedure, outlined in Fig. 1, is used to estimate phenomenological rate coefficients k(T,P) for a reaction network of
arbitrary connectivity and complexity based exclusively on
high-pressure-limit parameters. The parameters needed are:
The species in the network. For each species the thermodynamic parameters (DHof (298 K), DSof (298 K), and a model
for Cp(T)) are needed. For those species that are isomers in the
network, the chemical structure (connectivity only), molecular
weight, and Lennard-Jones parameters are also needed.
The reactions in the network. For each reaction the highpressure limit rate coefficient kN(T) is needed.
The bath gas. The molecular weight and the LennardJones parameters are required.
The collisional transfer probabilities model. Throughout
this paper the single exponential down model is used due to its
dependence on a single parameter; more complex models can
1134
Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
be used given sufficient information, but often this is not
available. Other inputs include the set of temperatures and
pressures at which to estimate the phenomenological rate
coefficients, and an interpolation model to be fitted to the
computed k(Tn,Pm) values to give smooth functions k(T,P).
All of the numerical methods used to estimate the phenomenological rate coefficients require discretization of the
energy domain into a vector E. Such a discretization is
characterized by a minimum energy Emin, maximum energy
Emax, and either a number of grains Ngrains or a energy spacing
DE. The minimum energy Emin is straightforward to determine: it is simply the ground-state energy of the lowest
molecular configuration on the potential energy surface.
(The ground-state energies on the potential energy surface
are estimated as the enthalpy at 0 K using the thermodynamic
parameters for each species.) The number of grains Ngrains or
energy spacing DE depend on the user’s desired level of
accuracy, and are left as parameters to be input.
However, the choice of Emax is not straightforward. A
choice of Emax that is too low may cause the accuracy of the
resulting rates to suffer due to neglect of important energies.
A choice of Emax that is too high may cause numerical effort to
be wasted on energies that do not contribute significantly;
worse, the master equation matrix becomes increasingly stiff
due the faster reaction rates observed at higher energies.
Nonetheless, an automatic procedure for determining Emax is
required for our desired application of automatic mechanism
generation. Our chosen procedure is as follows:
1. Guess an initial value Emax,0 that is a certain number of
kBT above the highest energy on the potential energy surface,
usually a transition state or a bimolecular reactant or product
channel.
2. Estimate the density of states and the equilibrium
distribution for the isomer with the highest ground-state
energy. For a calculation performed at multiple temperatures,
determine the equilibrium distribution at the maximum temperature, as this gives the widest distribution.
3. Determine the energy at which the tail of the equilibrium
distribution is some desired fraction of the maximum of the
distribution.
4. If a suitable energy is found, add to it the difference
between the highest energy on the potential energy surface and
the isomer’s ground-state energy. This is the maximum energy
Emax.
5. If no energy in the current range qualifies, repeat this
procedure starting from a higher value of Emax,0 until a suitable
Emax is found. In the interest of speed it is preferred to minimize
the number of times the density of states is estimated, so our
first choice of Emax,0 is deliberately chosen to be significantly
higher than the maximum energy in most cases.
Once a suitable set of energy grains has been selected, the
density of states r(E) is calculated for all isomers in the
network. For multiple temperature and pressure calculations,
this step need only be performed once. Two methods for
estimating the molecular degrees of freedom needed to compute r(E), the three-frequency and functional-group frequency
methods, are described in Sections 4.1 and 4.2, respectively.
The input includes a set of temperatures and pressures at
which to estimate the phenomenological rate coefficients.
This journal is
c
the Owner Societies 2012
Downloaded by Fritz Haber Institut der Max Planck Gesellschaft on 21 November 2012
Published on 07 December 2011 on http://pubs.rsc.org | doi:10.1039/C1CP22765C
View Article Online
Fig. 1
A general procedure for estimating pressure-dependent rate coefficients for reaction networks of arbitrary size and complexity.
There are certain steps that are only dependent on the current
temperature and not on pressure, so we place the iteration over
temperature as the outer loop and the iteration over pressure
as the inner loop. For example, the equilibrium distributions
are functions of temperature only and are calculated using the
formula
bi ðEr ; TÞ ¼ PN
ri ðEr ÞebEr
grains
s¼1
This journal is
c
ri ðEs ÞebEs
the Owner Societies 2012
ð9Þ
for isomer i at energy Er for energy grain r, where ri(E) is the
vibrational-rotational density of states for the isomer.
The microcanonical rate coefficients k(E) can be computed
using a variety of methods. If reactant and transition state
energies, geometries, and vibrational frequencies are available
(e.g. from a quantum chemistry calculation), then the RiceRamsperger-Kassel-Marcus (RRKM) method gives excellent
results. However, such information is often expensive to
obtain and nontrivial to automate, since it requires the sum
Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
1135
View Article Online
Downloaded by Fritz Haber Institut der Max Planck Gesellschaft on 21 November 2012
Published on 07 December 2011 on http://pubs.rsc.org | doi:10.1039/C1CP22765C
of states for the transition state. Furthermore, anharmonicity
effects—such as hindered internal rotations—become increasingly
important for larger polyatomic molecules, and anharmonic
corrections are often necessary to achieve rate coefficients at
accuracies better than an order of magnitude. An alternative
approach for when such information is not available is
to utilize the inverse Laplace transform to transform the
canonical rate k(T) in the high-pressure limit into the microcanonical rate coefficient k(E). The relationship between these,
recognized by Slater,42 is
k(E)r(E) = L1[Q(b)kN(b)]
(10)
37
Exact formulas exist for simple Arrhenius kinetics and
modified Arrhenius kinetics kN(T) = ATn exp(Ea/RT) for
n 4 1/2.31 The inverse Laplace transform method assumes
that the given k(T) expression is valid over the temperature
range from zero to infinity, and that the activation energy Ea is
physically equivalent to the reaction barrier height E0, both of
which are nontrivial approximations. Note that, for simple
Arrhenius kinetics (n = 0) the quantum Rice-RamspergerKassel (QRRK) method is essentially the same as the inverse
Laplace transform method.
The RRKM or inverse Laplace transform method is used to
determine the microcanonical rate coefficient for the forward
reaction only. The reverse rate coefficient is determined from
detailed balance. For an isomerization reaction A - B the
detailed balance equation is
kf(E)rA(E) = kb(E)rB(E)
(11)
Constants of proportionality in the density of states
become unimportant, as they cancel when taking the ratio
r(E)/Q(b), e.g. when computing the equilibrium distribution
b(E,T). In Section 4 we will use this to include an arbitrary
active K-rotor in the density of states expression. Thus, our
implementation of the master equation uses eqn (13) and (14).
The collision model is made up of the collision frequencies
oi(T,P) and the collisional transfer probabilities model
Pi(E,E 0 , T). As is common within the master equation literature, here oi is calculated using the Lennard-Jones collision
model. This should be a reasonable choice for collisions
between polyatomic molecules and weak colliders such as
noble gases and diatomics, but less accurate for collisions
between two large polyatomics or two very polar molecules.43
The collisional transfer probabilities function Pi(E,E 0 , T)
comes in a variety of forms, all of which share the general
trend that a collision is more likely to cause a small transfer of
energy than a large one. The most common form of Pi(E,E 0 , T)
is the single exponential down model
Pi(E,E 0 , T) = Ci(E 0 , T)exp[ai(T)(E 0 E)] E o E 0 (15)
where the activating equivalent is determined from detailed
balance. C is determined by normalization of the probability
distribution. The lone parameter aI(T) is the inverse of hDEdi.
To be consistent with thermodynamics, detailed balance must
be satisfied:
ri ðE 0 Þ
E0 E
0
Pi ðE; E 0 ; TÞ EoE 0
exp
Pi ðE ; E; TÞ ¼
ri ðEÞ
kB T
ð16Þ
and for a dissociation reaction A - B + C the equation is
kf(E)rA(E) = kb(E)rBC(E)
(12)
where rBC(E) is the convolved vibrational-rotational density
of states for species B and C, and also includes relative
translational motion. An alternative formulation incorporates
the macroscopic equilibrium coefficient Keq(T) and equilibrium
distributions bi(E,T) at each temperature. For isomerization
and dissociation, respectively, the alternative formulation is
kf(E)bA(E,T) = Keq(T)kb(E)bB(E,T)
(13)
kf(E)bA(E,T) = Keq(T)kb(E)bBC(E,T)
(14)
where bBC(E,T) is the combined equilibrium distribution for
species B and C. These two formulations are equivalent;
however, there are multiple reasons to use eqn (13) and (14)
instead of eqn (11) and (12):
Only the density of states of the unimolecular isomers
need be computed. This is a result of the assumption of
thermalized bimolecular channels, which means that we only
need to compute the product kb(E) bBC(E,T), and not the
individual values of kb(E) and bBC(E,T). (In eqn (4) the
required products are the Fimbm terms).
Only the reactive vibrational-rotational modes need be
included in the density of states. Missing modes will not affect
the observed equilibrium because we are imposing the macroscopic equilibrium via Keq(T). This is particularly important in
automatic mechanism generation, where we do not yet have an
efficient, accurate method of estimating the external rotational
constants.
1136
Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
At this point we have all of the information needed to
construct the full master equation matrix, and we are ready
to extract the phenomenological rate coefficients. Three methods
to do this, the modified strong collision (MSC), reservoir state
(RS), and chemically-significant eigenvalue (CSE) methods, are
described in detail in Sections 5.1, 5.2, and 5.3, respectively.
4
Methods for estimating the density of states
As described in Sections 1 and 2, formulation of the master
equation requires the density of states rvk for each isomer. The
relevant rovibrational degrees of freedom are those degrees of
freedom in which the energy of the excited species can be
randomized, i.e. the constant J and M modes are not included.
In RMG, these degrees of freedom are taken to be the
harmonic-oscillator (HO) vibrational frequencies, hindered
internal rotors, and a one-dimensional external rotation (i.e.
the K-rotor for a symmetric top approximation, which is
expected to strongly couple with vibrations via Coriolis terms).
Furthermore, it is assumed that the degrees of freedom are
independent. RMG assumes that each molecule may be
approximated as a symmetric top and that the density of
states for the one-dimensional K-rotor is:
1=2
1
1
rr ðEÞ ¼
ð17Þ
s jA BjE
where s is the rotational symmetry number for this rotor, A is
the unique rotational constant, and B% is the average of the two
This journal is
c
the Owner Societies 2012
View Article Online
Downloaded by Fritz Haber Institut der Max Planck Gesellschaft on 21 November 2012
Published on 07 December 2011 on http://pubs.rsc.org | doi:10.1039/C1CP22765C
most similar rotational constants. As stated in Section 3, since
% is a proportionality constant in the above expression,
s|A B|
we need not be concerned with its actual value, as it will cancel
when we evaluate equilibrium distributions. A discussion of
the accuracy of the K-rotor model is available from Zhu
et al.44
The density of states for hindered internal rotors is given by
the Laplace transform of the hindered-rotor partition function. RMG assumes a semi-classical approximation for the
hindered-rotor partition function:
QHRsemi ¼ QHRclassical
¼
V0 p
2s2 Br
QHOquantum
QHOclassical
1=2
eau
2a
e I0 ðuÞ
1 e2au
1=2 u
u
ð18Þ
where Br is the rotational constant corresponding to the
reduced moment of inertia, V0 is the rotational barrier height
assuming a single cosine potential, s is the symmetry number
of the internal rotor, n is the normal mode frequency corresponding to the torsion, I0 is the modified Bessel function of the
first kind, u V0/2kBT, and a n/V0. Like with the K-rotor,
the rotational constant Br appears only as a proportionality
constant, so its value will not be important for our master
equation calculation. If detailed information about the species
or transition state is available, we compute the reduced
moment of inertia using the method of Pitzer45 which is
described as I(2,3) by East and Radom.46
The density of states corresponding to the classical hindered
rotor partition function was described by Knyazev.47 Unfortunately there is no analytical expression for the density of
states corresponding to the semi-classical hindered rotor
partition function, eqn (18), since there is no analytical inverse
Laplace transform for the denominator in the quantum
harmonic oscillator partition function. However, the BeyerSwinehart method gives the exact density of states of the
quantum harmonic oscillator. By replacing this denominator
with a truncated Taylor expansion, the result from Knyazev
can be extended to include the density of states corresponding
to a semi-classical hindered rotor partition function:
8 pffiffiffiffiffiffiffiffiffi
~ 0
2K
E=V
>
<
pffiffiffiffiffiffiffiffi
ps
Br V0
ffi
rHRsemi ðEÞ ¼ 2K pffiffiffiffiffiffiffiffi
V0 =E~
>
:
pffiffiffiffiffiffi
ps
Br E~
for Eo max ðV0 ; hn=2Þ
otherwise
ð19Þ
Above, Ẽ E hn/2, and K is the complete elliptical integral
of the first kind. The density of states eqn (19) is calculated for
each hindered rotor. The combined density states for multiple
hindered internal rotors is calculated numerically by repeated
application of the convolution integral. Similarly, the density
of states for internal rotors and active external rotation is
calculated by convoluting the density of states for the active
K-rotor with the density of states for all the hindered rotors.
Finally, following the method of Astholz,20,34 the BeyerSwinehart algorithm18 is initialized with the K-rotor/
hindered-rotor density of states, which convolutes this density
of states with the vibrational modes. The result is the complete
rovibrational density of states, r(E).
This journal is
c
the Owner Societies 2012
Accurate determination of the vibrational frequencies and
the potentials for internal rotation requires a threedimensional geometry and a quantum calculation, which is
not routinely available in the RMG framework. Instead, these
parameters are estimated from the heat capacity. The heat
capacity may be separated into contributions from external
translation, external rotation, and the internal rovibrational
degrees of freedom:
Ctotal
= Ctranslation
+ Crotation
+ Crovibration
V
V
V
V
In this presentation, the K-rotor is included in the external
rotation. The vibrational frequencies and hindered rotor
parameters can be calculated by fitting a function to the
vibrational contribution to the heat capacity. In most cases,
however, the number of vibrational modes exceeds number of
known values of the heat capacity as a function of temperature,
and therefore these frequencies cannot be determined
uniquely. Two methods for approximating the vibrational
frequencies are described below.
4.1 The three frequency model
The three frequency model for estimating r(E) from heat
capacity data was proposed by Bozzelli, Chang, and Dean.33
This approach assumes that the molecule can be described by
the HO model. The heat capacity of a single mode as a
function of its HO frequency is given by the following formula
from statistical mechanics:
2
CVHO ðT; n i Þ
n i =kB T
¼ eni =kB T n =k T
ð20Þ
R
e i B 1
where ni is the vibrational frequency for the ith mode. Bozzelli
et al. fit three vibrational frequencies to the heat capacity at
seven temperatures (300, 400, 500, 600, 800, 1000, 1500 K),
using the following formula:
CVvibration ðTÞ ¼
3
X
i¼1
gi CVHO ðT; n i Þ
ð21Þ
g3 ¼ s g1 g2
where s is the number of vibrational modes (e.g. 3N 6 for a
nonlinear molecular with no internal rotors). In the original
implementation, the three degeneracies gi were allowed to be
non-integers. Later implementations of this approach forced
the degeneracies to be integers, so r(E) could be computed
efficiently using the Beyer-Swinehart algorithm, and that is what
we have done here.27 Thus, five parameters are fit to the heat
capacity data: three frequencies and two (integer) degeneracies.
This method is simple and computationally efficient. However,
there are two problems with this approach. First, the resulting
pseudo-frequencies may not be physically realistic. Second, it
assumes that the heat capacity necessarily increases monotonically with temperature. For molecules with hindered
internal rotors, the true heat capacity may actually have a
local maximum with respect to temperature.
4.2 The functional-group frequency model
An alternative to the three-frequency model, inspired by infrared spectroscopy, is to assign frequencies to functional groups.
Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
1137
Downloaded by Fritz Haber Institut der Max Planck Gesellschaft on 21 November 2012
Published on 07 December 2011 on http://pubs.rsc.org | doi:10.1039/C1CP22765C
View Article Online
For each functional group, the number of characteristic frequencies is limited to 3Natoms Nrotons 6. For example, a heavy
atom attached to a methyl group contains 5 atoms and one
single bond between heavy atoms, and thus has 9 characteristic
frequencies: 3 C–H stretches, 1 R–C stretch, 2 R–C–H bending
modes, 2 R–C–H rocking modes, and 1 umbrella mode. Additionally, one of the six subtracted modes is the overall rotation of
the methyl group, and this torsional mode is treated as a hindered
internal rotor. RMG automatically recognizes that single bonds
between heavy atoms that are not within a ring structure may be
hindered rotors. Each vibrational mode is associated with a range
of frequencies, and RMG automatically selects the appropriate
number of frequencies, taking them to be evenly spaced between
the two bounds. Thus, for a molecule with two methyl groups,
RMG would assume 6 C–H stretching frequencies that are evenly
spaced between 2750–2850 cm1. A list of the functional groups,
the corresponding range of frequencies, and the number of
frequencies per group are available in the Supporting Materials.w
By summing over the functional groups, the model predicts the
majority of the HO vibrational frequencies. The heat capacity for
the remaining degrees of freedom (e.g. vibrational frequencies not
explicitly accounted for, as well as hindered rotors) is obtained by
subtracting the heat capacity of the functional group frequencies
from the rovibrational heat capacity:
CVremaining degrees of freedom ðTÞ ¼ CVrovibration ðTÞ
Nfunctional group frequencies
X
i¼1
CVHO ðT; n i Þ
ð22Þ
The remaining vibrational frequencies and the parameters
for hindered internal rotors are determined from this heat
capacity data. Unlike the three frequency model, this method
includes a function for hindered internal rotors and therefore
can accommodate heat capacities with a local maximum.
Nremaining
CVremaining degrees of freedom
X
¼
i¼1
þ
NX
rotors
j¼1
CVHO ðT; n i Þ
CVHR ðT; V0;j ; n j Þ
where Nremaining 3N 6 Nrotors Nfunctional group frequencies.
The heat capacity of a hindered internal rotor is derived from
the semi-classical model for the hindered rotor partition
function, eqn (18). The heat capacity for this partition
function is:
d ln QHRsemi
CVHR ðT; V0 ; nÞ
d
¼
T2
dT
R
dT
d2 ln QHRsemi
du2
"
#
1
I1 ðuÞ
I1 ðuÞ 2
u
þu u
2
I0 ðuÞ
I0 ðuÞ
¼ u2
þ
1138
2uaeau
1 e2au
2
Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
ð23Þ
Although the functional group method will estimate most of
the frequencies, it may be the case that not all of the remaining
parameters can be determined uniquely from the heat capacity
data. Typically, the CV(T) data from a group additivity
calculation has enough information content to determine
about six parameters. If there are more than six unknown
parameters, then some of the rigid-rotor harmonic-oscillator
frequencies or the hindered rotor parameters (or in some cases
both) may be lumped into a single frequency or hindered
rotor, respectively, so that no more than six parameters are
fitted:
N
X
i¼1
M
X
j¼1
CVHO ðT; n i Þ ! NCVHO ðT; nÞ
CVHR ðT; V0;j ; n j Þ ! MCVHR ðT; V0 ; nÞ
Alternatively, one can sometimes obtain a better fit and/or fit
additional parameters directly by first fitting an interpolation
model to the known CV(T) points, and choosing a higher
density of points to use when performing the frequency fitting.
We use this approach with the acetyl + oxygen case study
discussed in Section 6, using the Wilhoit polynomial model48,49
to directly fit up to ten parameters.
In order to make the resulting parameters as realistic as
possible, the fitted parameters are given physically meaningful
constraints: the harmonic oscillator frequencies are bound
between 180 and 4000 cm1, the hindered rotor torsional
frequencies are bound between 40 and 600 cm1, and the
hindered rotor barrier heights are bound between 10 and
10 000 cm1. The FORTRAN 90 code DQED50,51 was used
for the bounded, constrained, non-linear least-squares fitting
procedure.
5 Methods for estimating phenomenological rate
coefficients
The objective of each of the methods described in the following sections is to transform the large master equation matrix
from eqn (4) into a small number of phenomenological rate
coefficients k(T,P). All of the methods share a common
formalism in that they seek to express the population distribution vector pi for each unimolecular isomer i as a linear
combination of the total populations xj(t) and ymA(t) ymB of
unimolecular isomers Cj and bimolecular reactant channels
Am + Bm:
pi ðtÞ ¼
N
isom
X
j¼1
xj ðtÞuij þ
N
reac
X
m¼1
ymA ðtÞymB vim
ð24Þ
The vectors uij represent the portion of the population
distribution of energy states of unimolecular isomer i that
tracks the population of isomer j. In the modified strong
collision and reservoir state methods, this is because the
energy levels of isomer i are in pseudo-steady-state relationships with isomer j. The interpretation is a bit different in
the chemically-significant eigenvalues approach, but the
form of the equations is the same. Similarly, the vectors
This journal is
c
the Owner Societies 2012
Downloaded by Fritz Haber Institut der Max Planck Gesellschaft on 21 November 2012
Published on 07 December 2011 on http://pubs.rsc.org | doi:10.1039/C1CP22765C
View Article Online
vim represent the population distribution of energy states
of unimolecular isomer i that tracks the population of
bimolecular reactant channel m. Both uij and vim are independent of time.
Notice that, even for the best uij and vim, eqn (24) is an
approximation to the true pi(t) given by the solution of the full
master equation. This approximation is a form of model
reduction, and is done so that we can use the conventional
phenomenological kinetics equations, and do not have to
keep track of the evolving energy distributions of each
species.
The phenomenological rate coefficient matrix can be
constructed from the full master equation matrix M
shown in eqn (4) and the population distribution vectors uij
and vim:
Ngrains
kij ðT; PÞ ¼
X
s¼1
ðMi uij Þs þ
N
grains
isom NX
X
‘ai
Ngrains
kim ðT; PÞ ¼
X
s¼1
ðMi vim Þs þ
s¼1
ðKi‘ u‘j Þs
N
grains
isom NX
X
‘ai
s¼1
ð26aÞ
ðKi‘ v‘m Þs
ð25bÞ
Ngrains
þ
X
s¼1
ð25aÞ
knm ðT; PÞ ¼
þ
N
isom
X
‘¼1
N
isom
X
‘¼1
gn‘ u‘j
ð25cÞ
gn‘ v‘m
n¼1
ð26bÞ
ynA ðtÞynB wn ðEÞ
31
As described by Chang, Bozzelli, and Dean, the modified
strong collision method utilizes a greatly simplified collision
model that allows for a decoupling of the energy grains. In the
simplified collision model, collisional stabilization of a reactive
isomer is treated as a single step process, ignoring the effects of
collisional energy redistribution within the reactive energy
space. An attempt to correct for the effect of collisional energy
redistribution is made by modifying the collision frequency
oi(T,P) with a collision efficiency bi(T), following early work
by Troe on the single-isomer ‘‘fall-off’’ case.30,52 By approximating the reactive populations as existing in pseudo-steady
state, a matrix equation is formed at each energy through
which uij and vim vectors, and subsequently the k(T,P) values,
can be determined.
the Owner Societies 2012
mi ðEÞ ¼ oi ðT; PÞbi ðTÞ
ð25dÞ
The modified strong collision method
c
N
reac
X
where xi(t) reflects the total population of the nonreactive
energies of isomer i, dij is the Kronecker delta, and
Above, the indices i and j represent unimolecular isomers
of the initial adduct, m represents bimolecular reactants,
n represents bimolecular reactants and products, and s represents an energy grain. Thus, the rate coefficients above are
for isomerization, association, dissociation, and bimolecular
reactions Am + Bm - An + Bn, respectively. The various
methods provide different values of uij and vim.
This journal is
N
isom
X
d
pðE; tÞ ¼ LðEÞpðE; tÞ þ
xi ðtÞzi ðEÞ
dt
i¼1
ðFim bm Þs
knj ðT; PÞ ¼
5.1
After applying the modified strong collision approximation,
a population balance performed at a reactive energy E gives
2
2
3
32
3
p1 ðE; tÞ
m1 ðEÞ k12 ðEÞ . . .
p1 ðE; tÞ
6
6
7
76
7
6
7
76
7
d6
6 p2 ðE; tÞ 7 ¼ 6 k21 ðEÞ m2 ðEÞ . . . 76 p2 ðE; tÞ 7
6
6
6
7
7
7
dt 4
4
5
54
5
..
..
..
..
..
.
.
.
.
.
2
3
oi ðT; PÞbi ðTÞbi ðE; TÞdi1
6
7
N
isom
X
6 o ðT; PÞb ðTÞb ðE; TÞd 7
i
i
i2 7
i
xi ðtÞ6
þ
6
7
4
5
i¼1
..
.
2
3
f1n ðEÞbn ðE; TÞ
6
7
N
reac
X
6 f ðEÞb ðE; TÞ 7
7
2n
n
ynA ðtÞynB 6
þ
6
7
4
5
n¼1
..
.
Nreac þNprod
X
n¼1
N
isom
X
jai
kji ðEÞ
ð27Þ
gni ðEÞ
is the total rate of loss from isomer i. Applying the pseudosteady state approximation to the above gives
LðEÞpðEÞ ¼
N
isom
X
i¼1
xi ðtÞzi ðEÞ
N
reac
X
n¼1
ynA ðtÞynB wn ðEÞ ð28Þ
Solving these gives
p(i)(E) = [L(E)]1zi(E)
p(n)(E) = [L(E)]1wn(E)
(29)
where p(i)(E) is the pseudo-steady population of each isomer at
reactive energy E resulting from the thermal activation of
isomer i, and p(n)(E) is the pseudo-steady population of each
isomer at reactive energy E resulting from the chemical
activation of reactant channel n. These can then be used to
compute the k(T,P) values:
Ngrains
kij ðT; PÞ ¼ oi ðT; PÞbi ðTÞ
knj ðT; PÞ ¼
N
grains
isom NX
X
i¼1 r¼r0;i
X
r¼r0;i
ðjÞ
pi ðEr Þ
ðjÞ
gni ðEÞpi ðEr Þ
ð30aÞ
ð30bÞ
Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
1139
View Article Online
Ngrains
kim ðT; PÞ ¼ oi ðT; PÞbi ðTÞ
Downloaded by Fritz Haber Institut der Max Planck Gesellschaft on 21 November 2012
Published on 07 December 2011 on http://pubs.rsc.org | doi:10.1039/C1CP22765C
knm ðT; PÞ ¼
N
grains
isom NX
X
j¼1 r¼r0;j
X
r¼r0;i
ðmÞ
pi ðEr Þ
ð30cÞ
ðmÞ
ð30dÞ
gnj ðEr Þpj ðEr Þ
The indices above have the same meaning as in eqn (25): i and j
represent unimolecular isomers, m represents bimolecular
reactants, n represents bimolecular reactants and products,
and r represents energy grains.
The modified strong collision method can be expressed in
terms of the notation presented in the introduction to this
section, eqn (24):
bi ðEr Þdij Er oEj;crit
uij ðEr Þ ¼
ð31aÞ
ðjÞ
pi ðEr Þ Er Ej;crit
vim ðEr Þ ¼
0
Er oEm;crit
ðmÞ
pi ðEr Þ Er Em;crit
ð31bÞ
where dij is the Kronecker delta. The critical energy that
divides the low-energy Boltzmann grains from the high-energy
steady-state grains is determined as the first reactive energy
grain, i.e. the lowest energy where any microcanonical rate
coefficient k(E) is nonzero. Inserting the above equations into
eqn (25) gives identical k(T,P) values as those in eqn (30c),
(30d), (30a) and (30b).
5.2
The reservoir state method
An alternative to approximating collisional stabilization as a
single-step process, proposed by Green and Bhatti,32 is to
assume that, except for a depleted region near the transition
state energy, the low energy grains are Boltzmann distributed.
Therefore, we partition the energy grains for each isomer; the
low-energy grains form the reservoir, while the high-energy
grains form the active-state. Green and Bhatti suggest the
cutoff be placed a few kBT below the lowest transition-state
energy adjacent to that isomer; our work suggests that a better
choice is simply to place the cutoff at the lowest adjacent
transition-state energy, as the intervening grains are better
modeled as Boltzmann distributed than as pseudo-steady.
(This choice only matters at temperatures where the equilibrium population of these grains is significant). Here we
neglect all reactions (including tunneling) from these reservoir
grains; if tunneling below the transition state energy is
important, one should set the reservoir cutoff a little lower.
As the development below implies, the mathematics of the
reservoir state method parallels that of the modified strong
collision method. The primary difference is that, for the
reservoir state method, we will be working with all high-energy
grains for all isomers at once, rather than being able to treat
each energy grain independently. A more detailed treatment,
including a demonstration that the resulting k(T,P) values are
in fact independent of the initial condition, is provided in the
Supporting Materials.w
First, a Boltzmann distribution bri is imposed on the reservoir grains pri for each isomer i, i.e.
pri
1140
=
xi(t)bri
Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
(32)
leaving a single time-dependent constant of proportionality
xi(t) which is related to the total population of isomer i. Note
that a significant fraction of the equilibrium population is in
PNres;i r
the activated grains at high temperatures, i.e. s¼1
ðbi Þs o1.
r
The quantity bi is not renormalized, however, in order to
ensure that the computed k(T,P) values satisfy macroscopic
equilibrium. With the reservoir approximation, the master
equation for the active-state grains pai for each isomer is
2 aa
2 a3
32 a 3
p1
p1
Ka12 . . .
M1
6
6 7
76 7
6 a
76 7
7
aa
d6
6 pa2 7 ¼ 6 K21 M2 . . . 76 pa2 7
6
76 7
7
dt 6
4
54 5
4 5
..
..
..
..
..
.
.
.
.
.
2 ar r 3
M1 b1 d1j
6
7
N
isom
X
6 Mar br d 7
2 2 2j 7
xj ðtÞ6
ð33aÞ
þ
6
7
4
5
j¼1
..
.
2 a a 3
F1m bm
6
7
N
reac
X
6 Fa ba 7
6
7
2m
m
ymA ðtÞymB 6
þ
7
4
5
m¼1
..
.
N
N
isom
reac
X
X
d a
xj ðtÞzj þ
ymA ðtÞymB wm
p ¼ Lpa þ
dt
j¼1
m¼1
ð33bÞ
As with the modified strong collision method, the pseudosteady state approximation is applied to the active-state grains
to give a matrix equation
Lpa ¼
N
isom
X
j¼1
xj ðtÞzj
N
reac
X
m¼1
ymA ðtÞymB wm
ð34Þ
for which the solution is
pa(j) = L1zj
pa(m) = L1wm
(35)
a(i)
where p is the pseudo-steady active-state population of each
isomer resulting from the thermal activation of isomer i, and
pa(n) is the pseudo-steady active-state population of each
isomer resulting from the chemical activation of reactant
channel n. These can then be used to compute the k(T,P)
values via
X
aðjÞ
kij ðT; PÞ ¼
ðMra
ð36aÞ
i pi Þr
r
knj ðT; PÞ ¼
kim ðT; PÞ ¼
N
isom
X
i¼1
X
knm ðT; PÞ ¼
r
aðjÞ
ð36bÞ
aðmÞ
ð36cÞ
gni pi
ðMra
i pi
N
isom
X
i¼1
This journal is
Þr
aðmÞ
gni pi
c
ð36dÞ
the Owner Societies 2012
Downloaded by Fritz Haber Institut der Max Planck Gesellschaft on 21 November 2012
Published on 07 December 2011 on http://pubs.rsc.org | doi:10.1039/C1CP22765C
View Article Online
where pa(j)
and pia(m) contain the components of pa(j) and pa(m)
i
relating to isomer i. Once again, indices i and j represent
unimolecular isomers, index m represents bimolecular reactants, index n represents bimolecular reactants and products,
and index r represents energy grains. Note that we have not
needed the common pseudo-first-order assumption on bimolecular reactants (i.e. ymA { ymB) to arrive at the above result.
However, the steady-state approximations used in both the
modified strong collision and reservoir state methods implicitly assume that both ymA and ymB are small enough that ymA
ymBFimbm is small.
The terms of the pseudo-steady state vector pai and the
reservoir populations bri can be used to construct the uij and
vim vectors for use in eqn (24):
uij ¼
dij brj
aðjÞ
pi
vim ¼
0
aðmÞ
pi
ð37Þ
Again, using the above expressions with eqn (25) gives identical k(T,P) values as those from eqn (36).
Finally, a note on efficient solving of eqn (34). Collisional
energy transfer models favor small transfers of energy over
large ones. For the single exponential down model, P(E,E 0 ) 0 exponentially as |E 0 E| - N. This suggests that the
probabilities are negligible for sufficiently large energy transfers, which in turn results in a matrix Pi that is strongly
banded. We can take advantage of this bandedness by using
an indexing scheme that iterates over energies as the outer loop
and isomers as the inner loop. The equations in this manuscript utilize the opposite indexing scheme for ease of understanding, but our software implementation of these equations
uses the more efficient indexing. A typical sparsity pattern for
an active-state matrix L using the two indexing schemes
is shown in Fig. 2 for a three-isomer network. The grainmajor indexing scheme, Fig. 2b, yields a matrix with a much
narrower bandwidth than the isomer-major indexing scheme,
Fig. 2a.
5.3
The chemically-significant eigenvalues method
The following description of the chemically-significant eigenvalues method is based on the works of Pilling and
Robertson53,54 and Miller and Klippenstein.22,55,56 First, the
eigenvalues and eigenvectors of the master equation matrix in
eqn (4), denoted here as M, are computed:
M = QLQ1
(38)
The eigenvalue decomposition is usually performed after
symmetrizing M, which ensures that the calculated eigenvalues
K are real. Since this is a physical system, the eigenvalues are
nonpositive, with one eigenvalue of exactly zero corresponding
to the equilibrium distribution. (If irreversible product
channels are included, however, the zero eigenvalue will not
be present).
If the system contains Nchem isomers and reactant channels,
then we expect the master equation matrix to consist of Nchem
chemically-significant eigenvalues (including the zero eigenvalue, if present). Under conditions of high pressure and low
temperature, these eigenvalues should be distinct and lower in
magnitude (i.e. less negative) than the other eigenvalues, which
This journal is
c
the Owner Societies 2012
correspond to internal energy relaxation. The distinctness
criterion is especially important: if the largest magnitude
chemical eigenvalue is too close to the smallest magnitude
internal energy eigenvalue, the calculated k(T,P) values will
be nonsensical. We only consider the case where Nchem
chemically-significant eigenmodes can be identified.
The separation of the chemical and internal energy eigenvalues means that we can select a time after the initial energy
modes have relaxed but before significant chemical transitions
have occurred. After this short-time period, the trajectory of
the solution will be contained within the manifold defined by
the chemically-significant eigenvectors. Therefore we ought
to be able to construct the k(T,P) values using only the
chemically-significant eigenpairs, independent of the initial
condition.
First we construct an Nchem Nchem matrix Z from the
chemically-significant eigenvectors by summing all terms in
each eigenvector for each isomer:
X
Qs‘ i ¼ 1; 2; . . . ; Nisom ; ‘ ¼ 1; 2; . . . ; Nchem ð39Þ
Zi‘ ¼
s2i
For reactant channels there is only one element in each
eigenvector for that channel, so it is transferred to Z as-is:
Znl = Qnl
n = 1,2,. . .,Nreac, l = 1,2,. . .,Nchem
(40)
Physically the matrix Z corresponds to the eigenvector matrix
obtained from diagonalizing the matrix of phenomenological
rate coefficients K, i.e. K = ZL 0 Z1. For systems satisfying the
assumed separation of timescales, this allows us to directly
express the phenomenological rate coefficients for all reactions
between isomers and reactant channels:
kij ðT; PÞ ¼
NX
chem
‘¼1
l‘ Zi‘ Z‘j1 i; j 2 1; 2; . . . ; Nisom þ Nreac ð41Þ
To obtain k(T,P) values for dissociation and bimolecular
reactions to a product channel (infinite sink), we will need
an additional Nprod Nchem matrix Y with elements
Ym‘ ¼
N
isom
X
i¼1
ðiÞ
gmi q‘
m ¼ 1; 2; . . . ; Nprod
ð42Þ
Above, ql(i) contains only the elements of chemically-significant
eigenvector ql involving isomer i. The k(T,P) values for
reactions to the product channels are then
kmj ðT; PÞ ¼
Ncse
X
‘¼1
Ym‘ Z‘j1 j ¼ 1; 2; . . . ; Nisom þ Nreac
ð43Þ
The time-independent population distribution vector sets uij
and vim can be constructed by combining the full and reduced
chemically-significant eigenvectors via
uij(Er) = (QchemZ1)(i,r),j
vim(Er) = (QchemZ1)(i,r),m
(44)
where Qchem is a rectangular matrix containing only the
chemically-significant eigenvalues, and the subscripts indicate
the row of QchemZ1 corresponding to isomer i at energy grain r,
and the column corresponding to isomer j or reactant
channel m. Note that, since the original master equation
Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
1141
Downloaded by Fritz Haber Institut der Max Planck Gesellschaft on 21 November 2012
Published on 07 December 2011 on http://pubs.rsc.org | doi:10.1039/C1CP22765C
View Article Online
Fig. 2 A typical sparsity pattern for an active-state matrix of a three-isomer network using (a) an isomer-major indexing scheme and (b) an energy
grain-major indexing scheme. The latter gives a significantly more banded matrix, which can be used to accelerate the reservoir state linear solve.
matrix M depends on ymB, in the eigenvalue method the
k(T,P), uij, and vim depend on ymB. However, in most cases
only small ymB are of interest, for which these dependencies are
very weak.
The only new assumption utilized by the chemicallysignificant eigenvalues method is that the chemical and internalenergy relaxation timescales are distinct. When this is the
case, as it often is at low and moderate temperatures and high
and moderate pressures, the k(T,P) values will accurately
reflect the system behavior at chemical timescales. For high
temperatures and/or low pressures, one or more chemical
timescales become indistinguishable from the internal-energy
relaxation timescales. To apply the chemically-significant
eigenvalues method in this case, it is necessary to lump the
configurations that participate in the fast chemical relaxations
together into a single pseudospecies. To our knowledge, no
one has yet automated this process; although it certainly
would be possible to do so, it is beyond the scope of this paper.
Finally, Miller and Klippenstein discuss two ways of
extracting the k(T,P) values from the chemically-significant
eigenpairs. The above discussion presents the so-called
‘‘long-time’’ method, which has two advantages: it numerically
performs better as a chemical eigenvalue approaches the
continuum of internal energy eigenvalues, and it makes clear
that the k(T,P) values are not dependent on the initial
populations. The other method, called the ‘‘initial-rate’’
method, gives identical results at almost all conditions, and
may be slightly faster to execute since it avoids the small
matrix inversion required by the long-time method. A brief
discussion of the initial-rate method is provided in the
Supporting Materials.w
6
Case study: the acetyl + oxygen system
As an application of the various methods for estimating the
density of states and the phenomenological rate coefficients,
we turn our attention to the chemically-activated reaction of
acetyl radical with oxygen. This system is important to atmospheric chemistry as a step in the conversion of acetaldehyde
to peroxyacetylnitrate (PAN), the latter of which is a
1142
Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
secondary air pollutant.57 It is also potentially important in
the ignition chemistry of ethanol. The potential energy surface
for this reaction was computed originally by Lee, Chen, and
Bozzelli.58 The values in Fig. 3 are updated results, described
below. Our updated values are consistent with those recently
published by Villano et al.59
Michael, Keil, and Klemm demonstrated that the acetyl
radical is a major product of the reaction of acetaldehyde with
OH radical, which preferentially abstracts the weakly-bonded
aldehydic hydrogen. They also observed pressure-dependent
regeneration of OH radical when acetyl radical is reacted with
oxygen, ranging from nearly complete regeneration at low
pressures to minimal regeneration at high pressures.60 Later
experimental efforts generally agree on the relevant pressure
range being about 0.001 to 1 bar at 300 K.61–65 Recent
experiments have confirmed that the a-lactone is the carboncontaining product associated with OH regeneration.66
On the theoretical side, this system was previously studied
by Lee, Chen, and Bozzelli using a three-frequency model to
estimate the density of states and the modified strong collision
method to estimate the phenomenological rate coefficients.
Their calculations significantly underpredicted the regeneration of the OH radical.58 More recently Maranzana, Barker,
and Tonachini studied the system using quantum chemistry
calculated modes as input to the density of states and the
MultiWell stochastic master equation solver. They allowed the
collisional energy transfer parameters to vary so as to match
the experimental data, and thus were able to match the
regeneration of OH.67
The calculations of Lee et al. predicted that the dominant
bimolecular product channel was a biradical + OH. Hou et al.
highlight the unlikeliness of this product after recomputing the
potential energy surface using the G3MP2 compound method.63
Based on their results, we do not consider the biradical as a stable
configuration, and do not include it in our potential energy
surface or subsequent calculations. Additionally, Hou et al.
found a saddle point geometry for the association reaction and
used a rigid-rotor harmonic oscillator model to compute the
association rate coefficient. Depending upon which method they
chose—MP2, CCSD, G3MP2, CASPT2—the barrier height
This journal is
c
the Owner Societies 2012
Downloaded by Fritz Haber Institut der Max Planck Gesellschaft on 21 November 2012
Published on 07 December 2011 on http://pubs.rsc.org | doi:10.1039/C1CP22765C
View Article Online
Fig. 3
The RQCISD(T)/CBS//B3LYP/6-311++G(d,p) potential energy surface for the acetyl oxidation network.
varied between +13 kcal mol1 and 1 kcal mol1, relative to
the reactants. Given that the C–O bond distance in the saddle
point was 2.5 Å, the CASPT2(7,7) DE0 of 1.7 kcal mol1 is the
most reliable. At these distances the modes between the two
reacting moieties will be highly anharmonic, and a rigid-rotor
harmonic oscillator model is ill-advised. Using DFT, we were
able to locate a saddle point for this reaction as well, but
there were four vibrational modes with frequencies less than
20 cm1. These coupled, anharmonic modes correlate to the
relative orientation and separation of the two fragments. A
better way to treat this entrance channel would be to use
variable reaction coordinate transition state theory.68–70
Ideally the interaction potential would be computed at the
CASPT2 level, with an active space of at least 7 electrons and
five orbitals—six electrons in four orbitals for O2, and one
electron in one orbital for the unpaired electron in acetyl, with
the possibility of adding additional electrons and orbitals to
account for the CQO p-bond and/or any oxygen lone pair
electrons in acetyl. However, these calculations are beyond the
scope of this work, and we have adopted the high-pressure
limit for acetyl + O2 recommended by Lee et al. for simplicity.
We performed our own RQCISD(T)/CBS//B3LYP/
6-311++G(d,p) calculations to generate the PES in Fig. 3
Table 1
Calculated high-pressure limit rate coefficients for acetyl + O2a
Reaction
A
b
Acetyl + O2 - acetylperoxy
Acetylperoxy - hydroperoxyl-vinoxy
Acetylperoxy - ketene + HO2
Hydroperoxyl-vinoxy - ketene + HO2
Hydroperoxyl-vinoxy - lactone + OH
4.4
2.3
2.6
5.3
1.9
12
10
109
109
1016
1017
n
Ea
0.0
0.75
1.2
1.0
1.1
0.0
23.2
34.1
29.5
27.2
The units for A are s1 for unimolecular reactions and in cm3
molecule1 s1 for bimolecular reactions, with Ea in kcal mol1. The
rate coefficient is k = A(T/1[K])nexp(Ea/RT). Computed from
RQCISD(T) variational TST calculations; see text for details. b Taken
from ref. 53.
a
This journal is
c
the Owner Societies 2012
using the methodology of Goldsmith et al.71 All DFT and
CBS-QB3 calculations were done using Gaussian03.72 All
MP2 and QCISD(T) calculations were done using MOLPRO.73
The high-pressure limit for each reaction rate coefficient except
acetyl + O2 was computed using variational transition state
theory, as implemented in Variflex.25 These results are provided in Table 1. Although these are high-level calculations,
it is doubtful that these numbers are the final word on this
important reaction system.
6.1 Density of states comparison
Fig. 4 shows the density of states, partition function, and heat
capacity for acetylperoxy and hydroperoxylvinoxy as calculated from vibrational and hindered rotor data estimated using
the three-frequency and group frequency approaches and
calculated from quantum chemistry, with the goal of judging
how the two approximate models compare to the detailed
quantum chemistry model. The parameters for the internal
modes as determined for each model are given in tables in the
Supporting Materials.w The quantum chemistry parameters
were obtained from geometry optimization and frequency
calculations using density functional theory with the B3LYP
functional and 6-311++G(d,p) basis set, as described in the
preceding section. The three-frequency and functional groupfrequency methods estimate the internal (vibrational and
hindered rotor) modes, and a K-rotor for the coupled external
rotational mode. For the purposes of this comparison, we do
the same with the quantum chemistry calculations.
Both acetylperoxy and hydroperoxylvinoxy have 18 internal
degrees of freedom to be represented by the three pseudofrequency-degeneracy pairs. Acetylperoxy contains 16 vibrational modes and two hindered rotor modes, while
hydroperoxylvinoxy contains 15 vibrations and three hindered
rotors. Nonetheless, the three frequency method will represent
all of these as three pseudofrequency-degeneracy pairs. As
shown in Fig. 4, we see that this method performs somewhat
Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
1143
Downloaded by Fritz Haber Institut der Max Planck Gesellschaft on 21 November 2012
Published on 07 December 2011 on http://pubs.rsc.org | doi:10.1039/C1CP22765C
View Article Online
Fig. 4 Comparison of densities of states, partition functions, and heat capacities for internal degrees of freedom for the (a) acetylperoxy and
(b) hydroperoxylvinoxy species using a three-frequency model (TF), group frequency model (GF), and quantum chemistry calculations (QC).
Both the TF and GF models reproduce the QC heat capacity data, but GF more consistently predicts the density of states than TF.
better for hydroperoxylvinoxy than for acetylperoxy. Our
quantum chemistry-calculated barrier heights for the
acetylperoxy hindered rotors are 1.22 and 6.12 kcal mol1,
while those for hydroperoxylvinoxy are 4.78, 6.26, and
14.7 kcal mol1. The high barriers to internal rotation make
the harmonic oscillator approximation fairly accurate for
hydroperoxylvinoxy.
For acetylperoxy, the group frequency approach provides
twelve of the eighteen internal modes as characteristic vibrational frequencies: nine from the CH3 group and three from
the COO group. For hydroperoxylvinoxy, eleven of the
eighteen internal modes are set via characteristic frequencies:
six from the CH2 group and five from the COOH group. In
both cases the remaining modes—four vibrations and two
hindered rotors for acetylperoxy, four vibrations and three
1144
Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
hindered rotors for hydroperoxylvinoxy—are few enough that
their parameters can be fit directly to the remaining heat
capacity data. The results in Fig. 4 show an improved estimate
of the density of states for both isomers when compared to the
three frequency method. In particular, the fitted hindered
rotor barrier heights for acetylperoxy using the group
frequency method are 0.99 and 4.68 kcal mol1, correctly
predicting a low-barrier and a moderate-barrier hindered
rotation.
6.2 Comparison of the phenomenological rate coefficient
methods
There are several means by which the three methods of
extracting phenomenological rate coefficients from the master
This journal is
c
the Owner Societies 2012
Downloaded by Fritz Haber Institut der Max Planck Gesellschaft on 21 November 2012
Published on 07 December 2011 on http://pubs.rsc.org | doi:10.1039/C1CP22765C
View Article Online
equation may be evaluated. Most pertinent is direct comparison of the k(T,P) values, as these are what we wish to use in
our kinetic models. The downside of this approach is that
there is no ‘‘true’’ set of k(T,P) values available to use as a
basis for comparison, since the solution of the master equation
does not exactly match the conventional reduced-order k(T,P)
model. To enable comparison with the true solution, we must
turn to comparison of concentration profiles, which we can
obtain by direct integration of the master equation. However,
we can go one step further by using eqn (24) to reconstruct the
approximate population distributions pi(E,t) for each isomer,
and compare those with one another and with the full master
equation solution. In this section we apply these methods of
comparison to the acetyl + oxygen reaction network.
6.2.1 Rate coefficients. The phenomenological rate coefficients for the acetyl + oxygen system were calculated by the
modified strong collision, reservoir state, and chemicallysignificant eigenvalue methods using the same r(E) and k(E)
from quantum chemical calculations, the same energy
discretization (constant 0.5 kcal mol1 grain spacing), and
hDEdowni = 200(T/300)0.85 cm1.74 Fig. 5 shows the k(T,P)
values for CH3CO + O2 - products at a variety of temperatures and pressures. (The equivalent values for acetylperoxy products and hydroperoxylvinoxy - products are available in
the Supporting Materials.w) As expected, at high pressures and
low temperatures, the dominant pathway is collisional stabilization to the acetylperoxy isomer. At low pressures, the
lactone + OH product channel is dominant at low and
moderate temperatures, while the ketene + HO2 channel is
dominant at high temperature. All of this is consistent with
our physical understanding of the system.
The modified strong collision, reservoir state, and chemicallysignificant eigenvalue methods give very similar values of
the rate coefficients at high and moderate pressures and at
low and moderate temperatures. However, there are several
regions of disagreement. The effects of temperature and
pressure on the agreement varies with the type of product.
For reaction to the isomers acetylperoxy and hydroperoxylvinoxy, there is good agreement between the methods at low
temperature and high pressure; increasing temperature and/or
decreasing pressure causes the methods to diverge somewhat.
For reactions to the bimolecular product channels, the reverse
effect is observed: all methods disagree somewhat at low
temperature and high pressure, while increasing temperature
and/or decreasing pressure causes the methods to converge.
Note that the methods always seem to show consensus on the
rate coefficients for the major pathways, and only disagree on
the minor channels. However, depending on what aspect of the
system is of physical interest, having a good value for the
minor rate constants can be important.
6.2.2 Concentration profiles. The most direct method of
evaluating the performance of the three methods is to use the
k(T,P) values to generate predicted concentration profiles for
each approximate method and compare them to those
produced by solving the full time-dependent master equation
with a stiff ODE solver. The objective here is to evaluate how
well the xi(t) and ymA(t) reflect the total populations of isomer i
This journal is
c
the Owner Societies 2012
and reactants/products m. For these calculations, the initial
concentrations were set to a small amount of acetyl in an
atmospheric ratio of oxygen and nitrogen such that the total
was consistent with the temperature and pressure chosen. The
lactone + OH and ketene + HO2 product channels were
treated as irreversible sinks. Both low-temperature (500 K)
and high-temperature (1500 K) conditions were studied; in
each case the pressure was chosen to be as low as possible
without obtaining unphysical results with the chemicallysignificant eigenvalue calculation.
Fig. 6 compares the concentration profiles from each
method with those from the full solution at the same sets of
conditions. Under all conditions shown, all three methods
exhibit a period at short times before their profiles match those
from the full master equation solution. This time period
reflects the time scale of collisions, and is generally of the
order of 10o1, where o is the mean collision frequency at the
conditions being studied. Once a suitable number of collisions
have occurred, all of the methods predict the total concentrations of reactants, products, and isomers very well. For the
modified strong collision and reservoir state methods, the poor
performance at short times comes from use of the pseudosteady state approximation, while for the chemically-significant
eigenvalues method, this comes from treating the internal
energy eigenmodes as completely relaxed.
The first set of conditions, 500 K and 0.1 bar, represent a
low temperature, moderately low pressure system. Under these
conditions acetylperoxy is the primary product, although the
pressure is low enough that a significant amount of lactone is
also produced. All of the methods approximate the total
concentrations of reactants, products, and isomers very well
after the short-time period of about 108 s. The modified
strong collision method does deviate from the others slightly
(most notably for hydroperoxylvinoxy), but remains within an
order of magnitude of the true profiles.
Also interesting is the effect of high temperature, demonstrated using conditions of 1500 K and 1 bar. This time the
primary product is the ketene channel, with relatively little of
each isomer produced along the way. All of the reactant and
product channels are well-approximated by all of the methods,
but show some disagreement for the total concentrations of
the isomers. Of the three methods, the chemically-significant
eigenvalues method is clearly the most accurate in predicting
the isomer profiles, while the modified strong collision and
reservoir state methods are only slightly worse.
6.2.3 Population distributions. The concentration profile
comparison is valuable because it allows for comparison of
the three methods with the ‘‘true’’ solution obtained from the
full master equation. However, the ‘‘true’’ solution provides
the full population distributions pi(E,t) for each isomer,
while the approximate solutions only provide the total
populations xi(t). In comparing concentration profiles, we
simply summed the population distributions to obtain xi(t),
but we can also do the reverse: reconstruct the approximate
pi(E,t) from xi(t) and ymA(t) ymB using eqn (24). Thus, we can
also examine how the approximate population distributions
evolve in time to see in more detail each method’s strengths
and weaknesses.
Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
1145
Downloaded by Fritz Haber Institut der Max Planck Gesellschaft on 21 November 2012
Published on 07 December 2011 on http://pubs.rsc.org | doi:10.1039/C1CP22765C
View Article Online
Fig. 5 Comparison of rate coefficients versus (a) pressure and (b) temperature for CH3CO + O2 - products estimated using the modified strong
collision (MSC), reservoir state (RS), and chemically-significant eigenvalue (CSE) methods. In the plots, A = acetylperoxy, B = hydroperoxylvinoxy, C = ketene + HO2, and D = lactone + OH. The error between the three methods is generally within an order of magnitude. However,
over the range where the CSE method can be applied, the RS method gives rate constants that better match the CSE values than the MSC method.
The CSE method runs into numerical difficulties at high temperature and low pressure, and at low temperature; see text for details.
Fig. 7 shows the approximate population distributions pi(E,t)
versus energy for the acetylperoxy isomer from each method and
the full solution at several times. Under both sets of conditions, the
high-energy grains of acetylperoxy are populated before collisional
stabilization begins to be important. The collisions tend to cause
only small changes in energy, so even after ten collision times there
have not been enough collisions to significantly populate the lowenergy grains. The result is that the low-energy grains are not yet
Boltzmann-distributed, and so none of the approximate methods
match the true populations at low energies. As time progresses to
100 and 1000 collsion times, the low-energy grains become
increasingly thermalized, and so the approximate methods much
more accurately reflect the full solution.
1146
Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
Of the three methods, the chemically-significant eigenvalue
method is the first to accurately match the entirety of the
‘‘true’’ population distribution. As the eigenvalue method
makes no assumption about the nature of the energies, it is
able to capture the non-Boltzmann, non-steady state behavior
of the high reservoir and near threshold grains (about 15 to
0 kcal mol1). (A side effect of this is that the eigenvalue
method predicts that the lowest-energy grains have slightly
negative populations at short times; this is why the distribution
drops sharply off at certain times.) The reservoir state method
distributions have (by assumption) a Boltzmann distribution
at low energies, which in reality is only established after many
collision times. However, only the most populated low-energy
This journal is
c
the Owner Societies 2012
Downloaded by Fritz Haber Institut der Max Planck Gesellschaft on 21 November 2012
Published on 07 December 2011 on http://pubs.rsc.org | doi:10.1039/C1CP22765C
View Article Online
Fig. 6 Concentration profiles at low-temperature and high-temperature conditions, comparing the modified strong collision (MSC), reservoir
state (RS), and chemically-significant eigenvalue (CSE) methods with the full master equation solution, for the acetyl + oxygen system as shown in
Fig. 3. In the plots, A = acetylperoxy, B = hydroperoxylvinoxy, C = ketene + HO2, D = lactone + OH, and E = acetyl + O2. The reservoir
state and chemically-significant eigenvalue methods both perform very well at low temperatures. All three methods are comparably accurate at
high temperatures even though they disagree with the full solution at short times. The vertical line indicates the average time between collisions
with the bath gas.
grains need to be Boltzmann-distributed before the total
population is predicted accurately. The modified strong collision
method population distributions are quite similar to the
reservoir state method distributions despite the significantly
different collision models used. The main discrepancy is in the
low reactive energies (around 6 to 0 kcal mol1 below the
reactant energy, but above the lowest isomerization barrier),
which the reservoir state method predicts more accurately
since it allows for collisional energy transfer between grains.
From the plots in Fig. 6 and 7, we see that the reservoir state
method performs most consistently at low temperatures, and is
still competitive at moderately high temperatures. The kinetics
of the major pathways are consistently predicted across all
methods; the disagreement appears only in the minor pathways. The chemically-significant eigenvalue method also
performs very well under the conditions shown—often equal
to or better than the reservoir state method—but would not be
successful at lower pressures due to the blending of the
chemical eigenmodes into the internal energy eigenmodes, as
shown in Fig. 8.
6.2.4 Eigenvalue analysis. In Fig. 5b the k(T,P) curves
given by the chemically-significant eigenvalues method appear
to stop as temperature decreases at all pressures. To examine
why, let us look in more detail at the eigenvalues themselves.
Treating the two bimolecular product channels as irreversible
sinks, there are three chemically-significant eigenvalues, and
no zero eigenvalue. The computed eigenvalues are plotted in
Fig. 8a and b as a function of temperature and pressure,
respectively. The figures show the four lowest-magnitude
eigenvalues, with l4 representing the lowest internal-energy
eigenmode.
In Fig. 8b the magnitudes of the chemical and internalenergy eigenvalues become increasingly disparate as temperature
This journal is
c
the Owner Societies 2012
decreases, indicating that the master equation matrix is
becoming increasingly stiff. Once the temperature becomes
too low, the large separation in time scales makes the lowmagnitude eigenvalues numerically difficult to calculate when
using double-precision arithmetic—as was used throughout
this work—resulting in one or more unphysical positive
eigenvalues.75,76 A quadruple-precision eigenvalue algorithm
would allow for successful determination of the eigenvalues at
low temperature, but we believe this to be unnecessary due to
the success of the reservoir state method at low temperatures.
The reservoir state method is not immune to numerical issues
due to stiffness, but this is generally only problematic for much
larger reaction networks than acetyl + oxygen.
6.3 Perturbations to the acetyl + oxygen potential energy surface
Although there are a few regions in Fig. 5 where the three
approximate methods for estimating k(T,P) values disagree,
the magnitude of the disagreement is typically within an order
of magnitude. In order to exaggerate the circumstances under
which one or more methods differ significantly, we introduce
one or more artificial perturbations to the potential energy
surface. These perturbations involve raising the ground-state
energies or lowering the transition-state barrier heights for one
or more configurations on the surface, in order to vary the
degree of chemical activation and the relative rates of isomerization. Several such perturbations will be discussed in what
follows. For all perturbations we have provided plots of
k(T,P) versus pressure for rate coefficients corresponding to
acetyl + oxygen as the reactants. Many other plots of k(T,P)
values and some concentration profiles are available in the
Supporting Materials.w
6.3.1 Acetyl + 20 kcal mol1. We begin with the case
where the acetyl + oxygen ground-state energy is artifically
Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
1147
Downloaded by Fritz Haber Institut der Max Planck Gesellschaft on 21 November 2012
Published on 07 December 2011 on http://pubs.rsc.org | doi:10.1039/C1CP22765C
View Article Online
Fig. 7 Population distributions for acetylperoxy versus energy after 10, 100, and 1000 collision times, comparing the modified strong collision
(MSC), reservoir state (RS), and chemically-significant eigenvalue (CSE) methods with the full master equation solution. At the conditions used
above, o1 is equal to 3.4 109 and 6.9 1011 s, respectively. The short-lived, high-energy reactive states immediately reach a steady state
which is accurately replicated by all three methods. After 100–1000 collision times, each method provides a reasonable approximation of the
population distribution of the full solution. The CSE method is the most accurate, while the RS and particularly MSC methods underpredict
the population of the adduct states with energies just below the reactant energy (6 to 0 kcal mol1) at lower temperatures. The zero of energy is
acetyl + O2. The arrows indicate the energy of the lowest transition state, which is the top of the ‘‘reservoir’’ in the RS method.
increased by 20 kcal mol1. This perturbation results in a
system that is much more chemically activated than before by
creating a significant range of energies below the ground state
of acetyl + oxygen but well above all of the transition
barriers. The potential energy surface now reflects a very
chemically-activated system, as the entire acetyl + oxygen
configuration is many hDEdowni above all other reaction
barriers in the system. This means that an excited acetylperoxy
adduct can survive several collisional stabilization events
and remain reactive. Thus we are not surprised to see that
very high pressure is required to trap the adducts, as seen in
Fig. 9a.
However, only the reservoir state and chemically-significant
eigenvalue methods capture this effect, while the modified
strong collision method does not. The modified strong collision method replaces the collision model from the full master
equation with a greatly simplified single-step activation/
deactivation process. This is not a great model of this perturbed
1148
Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
system, as the excited adduct can be stabilized by many
kcal mol1 without losing reactivity. The collision efficiency
b(T) is intended as a correction to improve the single-step
approximation; this perturbation demonstrates a situation
where b(T) does not provide an adequate correction. Concentration profile comparisons at both low and high temperature
conditions clearly show the modified strong collision method
provides a significantly less accurate prediction for the isomer
profiles, while the other methods perform noticeably better,
especially at low temperatures.
6.3.2 Acetylperoxy + 20 kcal mol1. Next we perturb the
system by artificially raising the ground-state energy of the
initial adduct acetylperoxy by 20 kcal mol1, creating a much
shallower well on the potential energy surface with a much
smaller density of states, and so faster k(E). This has the effect
of pushing more of the Boltzmann distribution of acetylperoxy
into the higher-energy reactive grains, and also encourages
This journal is
c
the Owner Societies 2012
Downloaded by Fritz Haber Institut der Max Planck Gesellschaft on 21 November 2012
Published on 07 December 2011 on http://pubs.rsc.org | doi:10.1039/C1CP22765C
View Article Online
Fig. 8 Eigenvalues of the master equation matrix at (a) several pressures at 500 and 1000 K and (b) several temperatures at 0.01 and 1 bar. As the
pressure is decreased or the temperature is increased, one or more of the chemical eigenvalues (l1 to l3) become indistinguishable from the internal
energy eigenvalues (l4). At low temperatures the eigensolver gives unphysical positive eigenvalues; the last correctly computed low-magnitude
eigenvalues are circled.
chemically-activated well skipping from acetyl + oxygen to
hydroperoxylvinoxy. In Fig. 9b we see that the k(T,P) values
from acetyl + oxygen to hydroperoxylvinoxy have increased as
a result of this perturbation. At 500 K, this increase is so
significant that hydroperoxylvinoxy is predicted to be the
dominant product over the range of 1 to 100 bar. All three
methods predict this behavior, although a chemical eigenvalue
(equilibration of the isomers acetylperoxy and hydroperoxylvinoxy) is often blended into the collisional energy relaxation
eigenspectrum, so the chemically-significant eigenvalue method
does not succeed with three chemical eigenvalues. For situations
like this, Miller and Klippenstein recommend treating the two
equilibrated isomers as a single pseudo-species, which would
correspond to two chemical eigenvalues.22
When concentration profiles are compared under low
pressure conditions—where the k(T,P) value for acetyl +
oxygen - acetylperoxy has fallen off—there is some disagreement between the methods on the acetylperoxy profile. This is
shown at both low-temperature and high-temperature conditions in Fig. 10.
At low temperature, the modified strong collision method
consistently overpredicts the acetylperoxy concentration, while
This journal is
c
the Owner Societies 2012
the reservoir state method consistently underpredicts it.
The chemically-significant eigenvalue method is significantly
more accurate than both of the other two approximate methods
in this instance. However, the modified strong collision and
reservoir state methods provide very good profiles for the other
species as well, with the reservoir state profiles slightly more
accurate. Furthermore, under these conditions acetylperoxy is a
minor species, so more error in the observed concentration may
be acceptable depending on the application. At high temperature the primary disagreement between the modified strong
collision and reservoir state profiles is again for acetylperoxy. It
is unsurprising that the reservoir state differs from the full
master equation solution, especially at short times, as a significant fraction of the Boltzmann population of acetylperoxy
now resides in the reactive grains. What is more surprising is
how accurate the modified strong collision method is for the
same conditions. The single-step collision model used in the
modified strong collision method is perhaps more accurate
when there are fewer nonreactive energies, especially when
coupled with an accurate collision efficiency b(T).
For this perturbed potential energy surface, the highpressure-limit rate coefficients kN(T) for reactions involving
Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
1149
Downloaded by Fritz Haber Institut der Max Planck Gesellschaft on 21 November 2012
Published on 07 December 2011 on http://pubs.rsc.org | doi:10.1039/C1CP22765C
View Article Online
Fig. 9 Comparison of rate coefficients versus pressure for CH3CO + O2 - products estimated using the modified strong collision (MSC),
reservoir state (RS), and chemically-significant eigenvalue (CSE) methods. In the plots, A = acetylperoxy, B = hydroperoxylvinoxy, C =
ketene + HO2, and D = lactone + OH. (a) Increasing the ground-state energy of acetyl + oxygen by 20 kcal mol1 causes the MSC method to
disagree more with the RS and CSE methods. (b) Increasing the ground-state energy of acetylperoxy by 20 kcal mol1 leads to significant
discrepancies between the methods, particularly at high temperature and high pressure. The CSE method fails to distinguish three chemical
eigenvalues at many of these conditions.
acetylperoxy as the reactant are the largest. At 500 K the
kN(T) values for acetylperoxy - hydroperoxylvinoxy, acetylperoxy - acetyl + O2, and acetylperoxy - ketene + HO2
are 9.7 109 s1, 3.4 108 s1 and 2.0 106 s1, respectively.
At the pressure of 10 bar used for the profiles in Fig. 10, these
are all smaller than the average collision frequency of o =
7.1 1010 s1. This implies that the collision and reaction
processes do indeed occur at different timescales, and the
chemically-significant eigenvalue method is able to resolve
three chemical eigenvalues. However, at 1500 K the kN(T)
values for the same reactions are 1.9 1011 s1, 1.5 1012 s1
and 2.0 1011 s1, respectively, which are all well above the
1150
Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
average collision frequency of o = 3.2 1010 s1 at 10 bar.
This suggests that the reaction time scale blends into the
collision time scale, and thus we are not surprised that the
chemically-significant eigenvalues method is unable to resolve
three chemical eigenvalues at these conditions.
6.3.3 Hydroperoxylvinoxy + 20 kcal mol1. A related
perturbation is to artificially raise the ground-state energy of
the other isomer, hydroperoxylvinoxy, by 20 kcal mol1, again
creating a much shallower well on the potential energy surface.
Similar to the last perturbation, this reduces the density of
states r(E), increases k(E), and causes more Boltzmann
This journal is
c
the Owner Societies 2012
Downloaded by Fritz Haber Institut der Max Planck Gesellschaft on 21 November 2012
Published on 07 December 2011 on http://pubs.rsc.org | doi:10.1039/C1CP22765C
View Article Online
Fig. 10 Concentration profiles at low-temperature and high-temperature conditions, comparing the modified strong collision (MSC), reservoir
state (RS), and chemically-significant eigenvalue (CSE) methods with the full master equation solution for the perturbed acetyl + oxygen system
where the ground-state energy of acetylperoxy is artificially increased by 20 kcal mol1. In the plots, A = acetylperoxy, B = hydroperoxylvinoxy,
C = ketene + HO2, D = lactone + OH, and E = acetyl + O2. At low temperature, all three methods agree with each other and the full master
equation solution except for acetylperoxy at short times, for which the CSE method provides the best match of the full solution. At high
temperature, the MSC profiles are noticeably more consistent with the full master equation solution than the RS profiles, particularly for
acetylperoxy. The CSE method is unable to find three chemical eigenvalues at the high temperature conditions, so its profiles are not shown).
population of hydroperoxylvinoxy to exist in the reactive grains.
Overall this encourages well skipping from acetylperoxy to the
product channels, and reduces the lifetime of the hydroperoxylvinoxy. The result is the k(T,P) values for reactions that
produce hydroperoxylvinoxy are depressed, as seen in
Fig. 11a. The only major disagreement in k(T,P) values
between the modified strong collision and reservoir state
methods are for those involving hydroperoxylvinoxy as a
reactant or product. (Again, the chemically-significant eigenvalue method for Nchem = 3 has no results because a chemical
eigenvalue has merged into the internal energy relaxation
eigenspectrum). Concentration profile comparisons to full
solutions of the master equation again show that the reservoir
state method performs slightly better under most conditions
for all configurations except the very unstable species hydroperoxylvinoxy, for which the modified strong collision method
is slightly more accurate.
6.3.4 Isomerization 20 kcal mol1. As a final perturbation, we lower the reaction barrier for the isomerization
reaction by 20 kcal mol1. The results of this perturbation
are shown in Fig. 11b. The low barrier causes the isomerization process to be very fast, to the point where a chemical
eigenvalue blends into the internal energy relaxation eigenvalues at nearly all conditions shown, causing the chemicallysignificant eigenvalues method to have no result with Nchem = 3.
The modified strong collision and reservoir state methods are
competitive, although at high temperatures the former performs noticeably better than the latter in predicting both
isomer profiles (figure available in the Supporting Materialsw).
6.3.5 Discussion. A few common themes emerge from the
above analysis. First, all three methods are fairly accurate for
This journal is
c
the Owner Societies 2012
the major channels across a wide range of conditions. Second,
the modified strong collision method performs consistently in
all systems, nearly always giving reasonable results, although
sometimes noticeably less accurate than the other methods.
The modified strong collision method is also numerically
inexpensive and robust. For initial explorations of a system,
the modified strong collision method is a reasonable choice.
Third, there is a wide range of conditions, especially at low
and moderate temperatures, where the reservoir state method
provides better estimates for the phenomenological rate coefficients than the modified strong collision method. Under these
conditions, the equilibrium population distributions for each
isomer are narrow enough that only a small fraction of the
Boltzmann population exists in reactive grains. When the
ground-state energies of either isomer were raised, the reservoir state method provided less accurate estimates for the
k(T,P) values involving the unstable isomer, under conditions
where the unstable isomer is a minor channel. When the
isomerization barrier was lowered, reservoir state method
estimates involving both isomers became less accurate. Thus,
we conclude that the reservoir state method is most accurate when
the isomer wells are ‘‘deep’’, i.e. the fraction of the Boltzmann
distribution that exists in reactive grains is insignificant.
Fourth, at conditions where the chemically-significant
eigenvalues method was able to resolve three chemical eigenvalues, it was consistently the most accurate in its k(T,P)
estimates. This is the best method to use for high-accuracy
k(T,P) calculations. However, there are a wide range of
conditions where this method was unsuccessful due to one
or more chemical eigenvalues being merged with the internalenergy relaxation eigenvalues, or because the numerical
eigensolver returned an unphysical positive eigenvalue. The
eigenvalue-merging issue, i.e. the breakdown of the assumed
Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
1151
Downloaded by Fritz Haber Institut der Max Planck Gesellschaft on 21 November 2012
Published on 07 December 2011 on http://pubs.rsc.org | doi:10.1039/C1CP22765C
View Article Online
Fig. 11 Comparison of rate coefficients versus pressure for CH3CO + O2 - products estimated using the modified strong collision (MSC),
reservoir state (RS), and chemically-significant eigenvalue (CSE) methods. In the plots, A = acetylperoxy, B = hydroperoxylvinoxy, C = ketene +
HO2, and D = lactone + OH. (a) Increasing the ground-state energy of hydroperoxylvinoxy by 20 kcal mol1 encourages well skipping between
acetylperoxy and the two product channels. The CSE method is unable to determine three chemical eigenvalues at all conditions due to the very
fast equilibrium between the isomers. (b) Decreasing the isomerization barrier height by 20 kcal mol1 creates a situation of a fast equilibrium
between the two isomers except at extreme pressure and low temperature, causing the CSE method to be unable to resolve three chemical
eigenvalues. The discrepancies in predicted isomer yields are usually not important, since the thermal equilibrium of the isomers is fast
(B1010 s1). The discrepancies in predicted yields of the product channels (C and D) are more important; the RS method is more accurate at low
temperature, while the MSC method is more accurate at high temperature.
separation of collision and reaction timescales, has been
discussed by Miller and Klippenstein; they suggest that under
these conditions, the concept of a phenomenological rate
coefficient breaks down.22 Unfortunately, both issues arise
frequently even for networks of modest size.
6.4
Computational cost considerations
Our results show that the reservoir state and chemicallysignificant eigenvalues methods are somewhat more accurate
1152
Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
than the modified strong collision method in predicting rate
coefficients for minor channels. The chemically-significant
eigenvalues method has a stronger theoretical basis, but the
reservoir state method is nearly as accurate.
An estimate of the relative execution time of the methods,
shown in Table 2, comes from considering the time needed by
each method to generate the data in Fig. 5. The reservoir state
method has about 2–3x the computational cost of the modified
strong collision method, while the cost of chemically-significant
This journal is
c
the Owner Societies 2012
View Article Online
Downloaded by Fritz Haber Institut der Max Planck Gesellschaft on 21 November 2012
Published on 07 December 2011 on http://pubs.rsc.org | doi:10.1039/C1CP22765C
Table 2
CPU time to generate Fig. 5
Method
Time (s)
Modified strong collision
Reservoir state
Chemically-significant eigenvalues
4.8
9.8
310
eigenvalues method is nearly 100x the modified strong collision method. These factors will only get larger as more isomer
wells are considered, or if finer energy discretization is
employed. The high computational expense and difficulties
inherent in the chemically-significant eigenvalues method at
low and high temperatures make it less useful for automatic
estimation of large numbers of k(T,P) parameters.
Given the approximate nature of several of the inputs to the
master equation—especially the collisional transfer probabilities model and parameters—there is at this point some
question as to whether the reservoir state method provides
enough of an improvement in accuracy compared with the
modified strong collision method to be worth the extra computational effort. This is especially true when we apply these
methods using automatically-generated estimated parameters,
which themselves can have a significant error. Furthermore,
the reservoir state method, while generally faster and more
robust than the chemically-significant eigenvalues method, is
not immune to numerical difficulties. The reservoir state
method involves a linear solve for the pseudo-steady activestate grain populations via Gaussian elimination. For very
large networks this elimination can be unsuccessful due to the
large matrices involved. We generally observe failures of the
LAPACK Gaussian elimination routines when there are
more than around thirty unimolecular isomers in the reaction
network (typically we allocate about 200 active-state grains
per isomer). For these reasons, our automatic mechanism
generation code, RMG, provides the user the option of either
the modified strong collision or reservoir state methods, and
automatically falls back to the modified strong collision
method if the reservoir state method is unable to complete
the Gaussian elimination step.
7
Conclusion
A general framework for estimating the phenomenological
rate coefficients for unimolecular reaction networks of arbitrary size and complexity using only macroscopic parameters
has been presented. Two important steps in the calculation
have been explored in detail: the modeling of the molecular
degrees of freedom used to determine the density of states, and
the methods for estimating the phenomenological rates from
the full master equation model. In the former, the functional
group frequency method has been shown to be generally
superior to the three frequency method, giving a density of
states that is more consistently close to that of the full
quantum chemistry-based model. In the latter, the reservoir
state method has been shown to be superior to the modified
strong collision method except at very high temperatures,
giving concentration profiles that more closely follow those
of the full master equation solution. These methods are
compared with the chemically-significant eigenvalues method
This journal is
c
the Owner Societies 2012
of Klippenstein and Miller and with the full master equation
solution in a way that we hope sheds some light on the
relationships between these methods.
All three approximations to the master equation perform
well at estimating the k(T,P) values of the major pathways of a
potential energy surface. However, for strongly chemicallyactivated systems, significant disagreement between the
methods is observed for minor pathways. The chemicallysignificant eigenvalues method is noticeably more computationally expensive than the other methods, but also gives the
most accurate estimates of the k(T,P) values. However, this
method has numerical difficulty at low temperatures due to
large separation of chemical and internal energy relaxation
timescales, and currently requires human adjustment to handle
fewer chemical eigenvalues for systems with a fast equilibration. This could perhaps be automated, but to our knowledge this automation is not currently available. The reservoir
state method is much less expensive than the chemicallysignificant eigenvalues method, and works very well at low
and moderate temperatures. However, when the reservoir
approximation is not valid—i.e. for high temperatures and
shallow wells—the reservoir state method gives worse estimates for the k(T,P) values than even the modified strong
collision method. This could perhaps be addressed by manually
removing the unimolecular wells which no longer correspond to stable species, but such a procedure has not yet been
made automated. The modified strong collision method is not
as accurate as the other methods in the ranges where those
methods work well—particularly for strongly chemicallyactivated potential energy surfaces with deep unimolecular
wells—but is computationally inexpensive and much more
robust over the full temperature range.
In practical applications of automatic mechanism generation,
most pressure-dependent reaction networks considered are very
complex, with a high likelihood of shallow wells being present.
For this reason, we recommend the modified strong collision
method for general use with automatic mechanism generation.
There is room for improvement by using the reservoir state or
chemically-significant eigenvalues methods in the ranges where
they are reliable, e.g. for reactions which the model is sensitive
to. In particular, the more rigorous methods could become
viable if one could automate removal of the very short-lived
wells at high temperature. However, one would need good error
control in the software to catch and recover from cases where
the more rigorous methods are not reliable.
The result of this work is a pair of modules for estimating
molecular degrees of freedom and phenomenological rates that
are utilized by and distributed with the open-source automatic
mechanism generation code RMG. RMG uses the functionalgroup frequency method exclusively and offers users a choice
between the modified strong collision and reservoir state
methods. Preliminary versions of both of these codes, written
in Fortran 90, are currently available in the RMG package.1
Acknowledgements
This work was funded in part by the US Department of
Energy, Office of Basic Energy Sciences, Division of Chemical
Sciences, Geosciences, and Biosciences, under Contract
Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
1153
View Article Online
DEFG02-98ER14914.; and by Award No. KUS-I1-010-01,
made by King Abdullah University of Science and Technology
(KAUST). Their financial support is gratefully acknowledged.
Downloaded by Fritz Haber Institut der Max Planck Gesellschaft on 21 November 2012
Published on 07 December 2011 on http://pubs.rsc.org | doi:10.1039/C1CP22765C
References
1 J. W. Allen, R. W. Ashcraft, G. J. Beran, C. F. Goldsmith,
M. R. Harper, A. Jalan, G. R. Magoon, D. M. Matheu,
S. S. Merchant, J. D. Mo, S. Petway, S. Ruman, S. Sharma,
K. M. Van Geem, J. Song, J. Wen, R. H. West, A. Wong,
H.-W. Wong, P. E. Yelvington, J. Yu and W. H. Green, RMG
(Reaction Mechanism Generator) version 3.2, 2010, http://rmg.
sourceforge.net/.
2 F. A. Lindemann, Trans. Faraday Soc., 1922, 17, 598–606.
3 C. N. Hinshelwood, Proc. Royal Soc. A, 1926, 17, 230–233.
4 O. K. Rice and H. C. Ramsperger, J. Am. Chem. Soc., 1927, 49,
1617–1629.
5 L. S. Kassel, J. Phys. Chem., 1928, 32, 1065–1079.
6 R. A. Marcus and O. K. Rice, J. Phys. Chem., 1951, 55, 894–908.
7 A. J. F. Siegert, Phys. Rev., 1949, 76, 1708–1714.
8 A. F. Bartholomay, Bull. Math. Biophys., 1958, 20, 175–190.
9 E. W. Montroll and K. E. Shuler, Adv. Chem. Phys., 1958, 1,
361–399.
10 I. M. Krieger and P. J. Gans, J. Chem. Phys., 1960, 32,
247–250.
11 P. J. Gans, J. Chem. Phys., 1960, 33, 691–694.
12 B. Widom, J. Chem. Phys., 1959, 31, 1387–1394.
13 J. Keck and G. Carrier, J. Chem. Phys., 1965, 43, 2284–2298.
14 J. Troe and H. G. Wagner, Ber. Bunsenges. Phys. Chem., 1967,
71, 937.
15 J. Troe, Ber. Bunsenges. Phys. Chem., 1973, 77, 665.
16 D. C. Tardy and B. S. Rabinovitch, Ber. Bunsenges. Phys. Chem.,
1973, 45, 3720–3730.
17 C. W. Gear, Commun. ACM, 1973, 14, 176–179.
18 T. Beyer and D. F. Swinehart, Commun. ACM, 1973, 16, 379.
19 S. E. Stein and B. S. Rabinovitch, J. Chem. Phys., 1973, 58,
2438–2444.
20 D. C. Astholz, J. Troe and W. Wieters, J. Chem. Phys., 1979, 70,
5107–5116.
21 J. R. Barker, Int. J. Chem. Kinet., 2001, 33, 232–245.
22 J. A. Miller and S. J. Klippenstein, J. Phys. Chem. A, 2006, 110,
10528–10544.
23 R. G. Gilbert, S. C. Smith and M. J. T. Jordan, UNIMOL program
suite (calculation of fall-off curves for unimolecular and recombination reactions), available from: School of Chemistry, Sydney
University, New South Wales, Australia, 1993.
24 V. Mokrushin, V. Bedanov, W. Tsang, M. R. Zachariah and
V. D. Knyazev, ChemRate Computer Program, version 1.10,
1999.
25 S. J. Klippenstein, A. F. Wagner, R. C. Dunbar, D. M. Wardlaw
and S. H. Robertson, Variflex Version 1.00, 1999, http://ftp.tcg.
anl.gov/pub/variflex/Summary.vrfx.
26 S. H. Robertson, D. R. Glowacki, C.-H. Liang, C. Morley and
M. J. Pilling, MESMER (Master Equation Solver for Multi-Energy
Well Reactions), 2008; an object oriented C++ program for
carrying out ME calculations and eigenvalue-eigenvector analysis
on arbitrary multiple well systems, 2008, http://sourceforge.net/
projects/mesmer.
27 D. M. Matheu, PhD thesis, Massachusetts Institute of Technology,
2003.
28 J. Troe, J. Phys.Chem., 1979, 83, 114–128.
29 J. Troe, Ber. Bunsenges. Phys. Chem., 1983, 87, 161–169.
30 R. G. Gilbert, K. Luther and J. Troe, Ber. Bunsenges. Phys. Chem.,
1983, 87, 169–177.
31 A. Y. Chang, J. W. Bozzelli and A. M. Dean, Z. Phys. Chem.,
2000, 214, 1533–1568.
32 N. J. B. Green and Z. A. Bhatti, Phys. Chem. Chem. Phys., 2007, 9,
4275–4290.
33 J. W. Bozzelli, A. Y. Chang and A. M. Dean, Int. J. Chem. Kinet.,
1999, 29, 161–170.
34 R. G. Gilbert and S. C. Smith, Theory of Unimolecular and
Recombination Reactions, Blackwell Scientific, Oxford, England,
1st edn, 1990.
1154
Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
35 T. Baer and W. L. Hase, Unimolecular Reaction Dynamics, Oxford
University Press, 1996.
36 K. A. Holbrook, M. J. Pilling and S. H. Robertson, Unimolecular
Reactions, John Wiley and Sons, 2nd edn, 1996.
37 W. Forst, Unimolecular Reactions: A Concise Introduction,
Cambridge University Press, 2003.
38 M. J. Pilling and S. H. Robertson, Annu. Rev. Phys. Chem., 2003,
54, 245–275.
39 S. H. Robertson, A. I. Shushin and D. M. Wardlaw, J. Chem.
Phys., 1993, 98, 8673–8679.
40 S. C. Smith and R. G. Gilbert, Int. J. Chem. Kinet., 1988, 20,
307–329.
41 J. A. Miller and S. J. Klippenstein, J. Phys. Chem. A, 2002, 106,
4904–4913.
42 N. B. Slater, Proc. Leeds Philos. Lit. Soc., 1960, 6, 691–694.
43 A. Fernández-Ramos, J. A. Miller, S. J. Klippenstein and
D. G. Truhlar, Chem. Rev., 2006, 106, 4518–4584.
44 L. Zhu, W. Chen, W. L. Hase and E. W. Kaiser, J. Phys. Chem.,
1993, 97, 311–322.
45 K. S. Pitzer, J. Chem. Phys., 1946, 14, 239–243.
46 A. L. L. East and L. Radom, J. Chem. Phys., 1997, 106, 6655–6674.
47 V. D. Knyazev, I. A. Dubinsky, I. R. Slagle and D. Gutman,
J. Phys. Chem., 1994, 98, 5279–5289.
48 R. C. Wilhoit, Thermodynamics Research Center Current Data
News, NIST, 1975, vol. 3, ch. 2.
49 W. C. Gardiner and A. Burcat, in Combustion Chemistry,
ed. W. C. Gardiner, Springer-Verlag, 1984, ch. 8.
50 R. Hanson and F. Krogh, DQED, http://www.netlib.org/opt/dqed.f.
51 J. Burkardt, DQED Fortran90, http://people.sc.fsu.edu/burkardt/
f_src/dqed/dqed.html.
52 J. Troe, J. Chem. Phys., 1977, 66, 4745–4757.
53 M. A. Blitz, K. J. Hughes, M. J. Pilling and S. H. Robertson,
J. Phys. Chem. A, 2006, 110, 2996–3009.
54 S. H. Robertson, M. J. Pilling, L. C. Jitariu and I. H. Hillier, Phys.
Chem. Chem. Phys., 2007, 9, 4085–4097.
55 S. J. Klippenstein and J. A. Miller, J. Phys. Chem. A, 2002, 106,
9267–9277.
56 J. A. Miller and S. J. Klippenstein, J. Phys. Chem. A, 2003, 107,
2680–2692.
57 J. H. Seinfeld and S. N. Pandis, Atmospheric Chemistry and
Physics–From Air Pollution to Climate Change, John Wiley and
Sons, 2nd edn, 2006.
58 J. Lee, C. J. Chen and J. W. Bozzelli, J. Phys. Chem. A, 2002, 106,
7155–7170.
59 S. M. Villano, N. Eyet, S. W. Wren, G. B. Ellison, V. M. Bierbaum
and W. C. Lineberger, J. Phys. Chem. A, 2010, 114, 191–200.
60 J. V. Michael, D. G. Keil and R. B. Klemm, J. Chem. Phys., 1985,
83, 1630–1636.
61 G. S. Tyndall, J. J. Orlando, T. J. Wallington and M. D. Hurley,
Int. J. Chem. Kinet., 1997, 29, 655–663.
62 M. A. Blitz, D. E. Heard and M. J. Pilling, Chem. Phys. Lett.,
2002, 365, 374–379.
63 H. Hou, A. Li, H. Y. Hu, Y. Li, H. Li and B. S. Wang, J. Chem.
Phys., 2005, 122, 224304.
64 R. K. Talukdar, M. E. Davis, L. Zhu and A. R. Ravishankara,
19th International Symposium on Gas Kinetics, 2006.
65 G. Kovács, J. Zádor, E. Farkas, R. Nádasdi, I. Szilágyi, S. Dóbé,
T. Bérces, F. Márta and G. Lendvay, Phys. Chem. Chem. Phys.,
2007, 9, 4142–4154.
66 S.-Y. Chen and Y.-P. Lee, J. Chem. Phys., 2010, 132, 114303.
67 A. Maranzana, J. R. Barker and G. Tonachini, Phys. Chem. Chem.
Phys., 2007, 9, 4129–4141.
68 Y. Georgievskii and S. J. Klippenstein, J. Phys. Chem. A, 2003,
107, 9776–9781.
69 S. J. Klippenstein, J. Chem. Phys., 1992, 96, 367–371.
70 S. J. Klippenstein, Y. Georgievskii and L. B. Harding, Phys. Chem.
Chem. Phys., 2006, 8, 1133–1147.
71 C. F. Goldsmith, S. J. Klippenstein and W. H. Green, Proc.
Combust. Inst., 2011, 33, 273–282.
72 M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria,
M. A. Robb, J. R. Cheeseman, J. J. A. Montgomery, T. Vreven,
K. N. Kudin, J. C. Burant, J. M. Millam, S. S. Iyengar, J. Tomasi,
V. Barone, B. Mennucci, M. Scalmani, G. Cossi, N. Rega,
G. A. Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota,
R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda,
This journal is
c
the Owner Societies 2012
View Article Online
73
74
75
76
W. Chen, M. W. Wong, C. Gonzalez and J. A. Pople, Gaussian 03,
2004.
H.-J. Werner, P. J. Knowles, R. Lindh, F. R. Manby and
M. Schuetz, et al., MOLPRO, version 2006.1, a package of ab
initio programs, 2006.
P. Zhang, S. J. Klippenstein, H. Sun and C. K. Law, Proc.
Combust. Inst., 2011, 33, 425–432.
T. J. Frankcombe and S. C. Smith, Comput. Phys. Commun., 2001,
141, 39–54.
T. J. Frankcombe and S. C. Smith, Faraday Discuss., 2001, 119,
159–171.
Downloaded by Fritz Haber Institut der Max Planck Gesellschaft on 21 November 2012
Published on 07 December 2011 on http://pubs.rsc.org | doi:10.1039/C1CP22765C
O. Kitao, H. Nakai, M. Klene, X. Li, J. E. Knox, H. P. Hratchian,
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, P. Y. Ayala, K. Morokuma, G. A. Voth,
P. Salvador, J. J. Dannenberg, V. G. Zakrzewski, S. Dapprich,
A. D. Daniels, M. C. Strain, O. Farkas, D. K. Malick,
A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. V. Ortiz,
Q. Cui, A. G. Baboul, S. Clifford, J. Cioslowski, B. B. Stefanov,
G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. L. Martin,
D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng,
A. Nanayakkara, M. Challacombe, P. M. W. Gill, B. Johnson,
This journal is
c
the Owner Societies 2012
Phys. Chem. Chem. Phys., 2012, 14, 1131–1155
1155