Você está na página 1de 50

View Article Online / Journal Homepage / Table of Contents for this issue

REVIEW www.rsc.org/annrepc | Annual Reports C

Product branching ratios in simple gas phase


reactions
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

Paul W. Seakins
DOI: 10.1039/b605650b

It has long been known that the product branching ratio between dierent
products in multichannel reactions is as important as the overall rate of
reaction, both in terms of practical applications (e.g. models of combustion
or atmospheric chemistry), and in understanding the fundamental mechan-
isms of such chemical reactions. However, it is only in the last 1015 years
that the experimental and theoretical tools have become available to
quantitatively address the issue of product branching ratios under relevant
conditions. This article includes a review of relevant developments in
experimental and theoretical techniques and then considers examples of
multichannel reactions. Firstly we examine the case of site specific reactions
taking place via the same basic mechanism, e.g. abstraction or addition and
then move on to review the class of multichannel reactions occurring with
the formation of a bound complex with subsequent unimolecular decom-
position to multiple sets of products. The final class of reactions considered
is where there is a competition between two dierent mechanisms e.g.
abstraction vs. addition. The review is concluded with a brief outlook on
to possible developments in the field.

1. Introduction
Simple gas phase reactions are the building blocks of complex chemical mechanisms
which control important phenomena such as combustion,1 plasmas,2 chemical
vapour deposition,3 atmospheric chemistry (of the Earth4 and other planetary
objects5,6) and the interstellar medium.7 A majority of reactions are bimolecular in
nature, i.e.
A+B-C+D (R 1)
and the rate of removal of reagents is given by:
dA#
! kbi A#B# 1
dt
where kbi is the bimolecular rate coecient for the reaction. Very often one of the
reagents is a radical or reactive species and the system can be experimentally
simplified by having the stable species in large excess minimising the potential for
radicalradical reactions. Such an approach also simplifies the solution of eqn (1); if
the molecular reagent, say B, is in great excess its concentration remains eectively
constant and kbi and [B] can be combined to give a pseudo-first-order rate coecient.

School of Chemistry, University of Leeds, Leeds, UK LS2 9JT

Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222 | 173


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

Eqn (1) now becomes:


dA#
! k0 A# 2
dt
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

where k 0 is the pseudo-first-order rate coecient, kbi ( [B]. Eqn (2) can be solved to
give the familiar exponential decay curve:
0
[A]t = [A]0e!k t (3)
where [A]t and [A]0 are the radical concentrations at time t and time zero,
respectively. A further major advantage of working under first or pseudo-first-order
conditions is that one does not need to measure the absolute concentration of the
radical species, but merely something that is proportional to concentration in order
to measure the overall rate coecient. This is especially important for radical species
where it is dicult to generate known absolute concentrations.
Rate coecients for a huge number of overall reactions determined by measuring
the removal of reagent species have now been made using a variety of dierent
methods to generate and monitor radical species. Databases (e.g. NIST, http://
kinetics.nist.gov/kinetics/) and critical evaluations (IUPAC http://www.iupac-
kinetic.ch.cam.ac.uk/ and JPL http://jpldataeval.jpl.nasa.gov/) are invaluable tools
for accessing this wealth of experimental information.
For simple reactions such as reaction (1) a measurement of the overall rate
coecient as a function of temperature may be all that is required, however many
radical reactions are more complex with several product channels being available. A
prime example is one of the most important reactions in combustion
k2a
H O2 !! OH O R2a

k2b
!! HO2 R2b

Reaction (2a) is chain branching (increasing the number of chain carrying


radicals) whereas reaction (2b) leads to the production of the relatively unreactive
HO2 radical and is eectively a chain termination reaction. Measurements on the
removal of H atoms will yield the overall rate coecient (k2 = k2a + k2b),
but no information on the relative yields of product. Clearly knowledge of the
fraction of the reaction proceeding down each channel is vital in combustion
modelling.
For a general bimolecular reaction
k3a
A B !! C D R3a

k3b
A B !! E F R3b

carried out under pseudo first order conditions where [B] c [A], the rate equation
for the formation of C is given by:
dC# 0 0
k03a A# k03a A#0 e!k3a k3b t 4
dt
solution of which yields:
A#0 k3a !fk0 k0 t
C#t e 3a 3b 5
k3a k3b

174 | Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

C is formed with the same time constant as A is removed and the final concentration
of C, [C]N, is given by:
A#0 k3a
C#1 6
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

k3a k3b
and the fraction of A proceeding via reaction (3a) is given by the product branching
ratio, BR:
k3a
BR3a 7
k3a k3b
As the specific rate coecients for each channel cannot be measured in simple kinetic
experiments, measurement of the branching ratio requires a measure of the final
product concentration, [C]p relative to the initial radical concentration [A]0. Many
experimental techniques for measuring radical concentrations are relative in nature
and it is obviously dicult to accurately generate a known radical concentration to
calibrate the measurements. In the review we outline dierent methods for measur-
ing branching ratios, but in essence they either require absolute concentration
measurements or having a calibration reaction which links the reagent and products
in a well defined way. A good example of the latter approach can be found in work
by Choi et al.8 on the H atom yield from the reaction of CN radicals with alkenes. A
calibration reaction (CN + H2) which proceeds 100% to form H atoms was used to
calibrate the H atom signal.
CN + H2 - HCN + H (R 4)
The H2 was then replaced with the alkene with all other conditions remaining
constant (i.e. the same CN concentration was generated) and the concentration of H
atoms was measured.
CN + C2H2 - C3H3N + H (R 4a)
- other products (R 4b)
Fig. 1 shows a schematic of the technique. Importantly the time dependence of the
CN removal and H atom production can be measured to demonstrate that the

Fig. 1 Schematic of calibration experiment for branching ratios. The solid line is the H atom
fluorescence from the calibration reaction, CN + H2 - HCN + H, with 100% conversion of
CN to H. The dashed line is the H atom fluorescence from the test reaction with an identical
initial concentration of CN radicals.

Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222 | 175


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

observed H is indeed coming from the target reaction and not any secondary or side
reactions.
Much work has been undertaken over the last 1015 years to improve the
experimental techniques allowing a greater number of species to be measured with
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

ever increasing sensitivity (increased sensitivity allows operation at lower radical


concentrations minimising secondary reactions) and we begin our review with a
discussion of recent advances in experimental techniques. The channel specific rate
coecients may not be able to be measured in simple kinetic experiments but they
can be calculated. One of the recent advances in this area has been the increased
accessibility of high level theoretical calculations and although the main focus of the
review is on experimental methods, a brief review of advances in theoretical
calculations are presented. This is a vibrant field of research and it is not possible
to do justice to all the work in the area. Hopefully this review will give a flavour of
the work that has taken place in the last ten years or so.
The review is broken down into various classes of reaction; site specific reactions,
reactions proceeding via complex formation and finally reactions proceeding via two
dierent mechanisms. Site specific reactions are the most straight forward and are
exemplified by the reaction of radicals (OH, Cl etc.) with alkanes, e.g. OH + C3H8.
The reaction can proceed via the abstraction of either a primary or secondary
hydrogen:
Cl + C3H8 - n-C3H7 + HCl (R 5a)
- i-C3H7 + HCl (R 5b)
Another example would be the addition of radical species to the double bond of an
un-symmetric alkene. The dominant class of reactions are those proceeding via
complex formation. The complex can be formed in a variety of waysradical
radical recombination, addition or insertion, and once the complex has been
formed it can rearrange or dissociate to a variety of dierent products. Multi-
mechanism reactions are those where the reaction can proceed via two dierent
mechanisms.

2. Recent developments in experimental and theoretical techniques


2.1 Extending the range of experimental measurements
Many of the applications of branching ratio studies can be found at extremes of
temperature. For example, very low temperatures (o20 K) for the interstellar
medium, low temperatures (80200 K) for the atmospheric chemistry of the Earths
upper atmosphere and the atmospheres of the outer planets to very high tempera-
tures (41700 K) for combustion.
The technique of laser flash photolysis (LFP) coupled with a range of detection
techniques has been the predominant method for kinetic and branching ratio studies
although discharge flow (DF) as a method of radical generation has some advan-
tages as well as some potential drawbacks. Flash photolysis has an advantage in that
it can operate over a wider range of conditions, typically 2001000 K and 2
1000 Torr can be routinely achieved and with specialist design these boundaries can
be significantly extended. For example, Hippler9 and Troe10 and their co-workers
have measured reactions at pressures up to 1000 bar and Fontijn11 developed a range
of high temperature reactors which closed the gap between conventional flash
photolysis and shock-tube measurements (typically greater than 1500 K). At the
other extreme of temperature Rowe, Smith and Sims have pioneered the use of Laval

176 | Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

expansions for studying radicalmolecule reactions from room temperature down to


1020 K.7,12,13 Recent developments include the introduction of pulsed Laval
nozzles1416 which reduce the extensive pumping capacity required in the original
Laval system at the expense of only being able to reach temperatures of B80 K.
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

One important issue regarding such expansions is the potential to generate clusters.
This can usually be avoided, but can be used to investigate the role of clusters in
catalysing reactions.17,18 To date Laval systems have mainly been limited to
monitoring radical removal rates via LIF, however, increasing expertise and the
coupling of a greater variety of detection techniques19 should soon lead to the
measurement of product branching ratios.
For the very high temperatures associated with combustion reactions shock tube
measurements still provide a majority of the experimental data. The main develop-
ments in this field have been to extend the range of real-time detection methods that
can be used to monitor the temporal profile of radical and molecular species. The
work of Hanson, Bowman and co-workers who have studied real-time profiles of a
variety of species including O, OH,20 CH321 and benzyl (C6H5CH2),22 are good
examples of the high quality data that can now be obtained. Fig. 2a shows an
example of an NH2 temporal profile obtained by frequency modulated absorption
spectroscopy23 and used to model the branching ratio for the NH2 + NO reaction
(see Section 4.2f). Careful design of experiment ensures that the measured profiles
are most sensitive to the target parameter as shown in Fig. 2b. Song et al.23 estimate
that the use of frequency modulation increases the detection sensitivity by a factor of
30 over direct absorption.
For very high temperatures thermal decomposition is generally used to generate
the radical species, however, at lower temperatures (13001700 K) it is possible to
shock to heat the system and then to initiate the reaction via laser photolysis of a
suitable precursor.24 Such techniques can have advantages in minimising the number
of radical species formed and hence simplifying the kinetic analysis.

2.2 Detection systems


Laser induced fluorescence (LIF) or resonance fluorescence has been the work-
horse radical and product detection system, oering excellent sensitivity, specificity
and selectivity. Excitation wavelengths between 600 and 200 nm can easily be
generated and LIF has been used to monitor the kinetics of a wide variety of radical
species and stable molecules. A recent development has been the use of frequency
tripling to extend the wavelengths of excitation light into the vacuum ultraviolet for
routinely monitoring species such as H,8 N25 and Cl.26 LIF is not a universal
technique requiring a bound accessible upper electronic state with a sucient
fluorescence quantum yield for detection and therefore development into other
techniques has continued.
Absorption spectroscopy has the potentially important advantage for branching
ratio studies in that, if the absorption cross section is known, intensity measurements
can be converted to absolute concentrations via the BeerLambert law:
IT = I0e!scl (8)
where IT and I0 are the transmitted and incident intensities, s is the absorption cross
section, c the concentration and l the path length. Measurements at radical
concentrations of typically less than 1013 cm!3 where the absorption will be small,
require measurements of two similar intensities; fluctuations in light output and
detector noise limit the sensitivity of such measurements. To a degree this can be

Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222 | 177


This journal is '
c The Royal Society of Chemistry 2007
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18. View Article Online

Fig. 2 (a) NH2 mole fraction profile from the NH2 + NO reaction at 1938 K and 1.19 bar.
The solid line is the best fit optimising the branching ratio, a = 0.63. The sensitivity of the fit to
variations in a is shown by the dashed lines. (b) Sensitivity analysis for the NH2 profile for the
conditions of Fig. 2a. Clearly the NH2 is most sensitive to the target parameter, a. Figures
reproduced with permission from ref. 23.

overcome by increasing the path length, an example being the elegant Herriot cell
design of Taatjes for IR measurements following laser flash photolysis.27
Selectivity in the UV region can be poor with many species having broad and
relatively featureless absorption spectra. The observed signal could be due to
contributions from several species with overlapping spectra. Diode array detectors,
as shown in Fig. 3 for studies on IO and BrO reactions, have allowed for the
collection of dispersed light as a function of time. Each row of the detector contains

178 | Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222


This journal is '
c The Royal Society of Chemistry 2007
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18. View Article Online

Fig. 3 (a) Schematic diagram of flash photolysis/UV diode array apparatus. (b) Dierential
absorption signal for BrO and IO. (c) Kinetic traces of BrO and IO. Figures reproduced with
permission from refs. 28 and 29.

a wavelength resolved spectrum (Fig. 3b) at a particular time following reaction


initiation.28,29 Such techniques have a considerable multiplex advantage over con-
ventional kinetic measurements which were obtained by fixing the monochromator
to a particular wavelength and monitoring the time dependence of the signal
following reaction initiation with the experiment subsequently repeated for dierent
wavelengths.
With its more structured spectra, the mid-IR region oers significant advantages
in terms of selectivity and specificity, however, absorption cross sections are
generally lower and high powered laser light sources are required. Unfortunately
tuneable lasers in the mid-IR have been dicult to operate although some success
has been achieved with F-centre lasers, particularly the work of Curl and co-
workers,30 and Leone and co-workers,31 primarily on straight kinetics (but with
some product studies32). Tuneable diode lasers oer more flexibility and there has
been more work on product branching ratios.3335 Quantum cascade lasers appear to
have the potential to provide better light sources for this region.36
An alternative method of generating tuneable mid-IR light is dierence frequency
mixing. The technique has been used by a number of groups.37,38 A further
alternative is to move away from the mid-IR and utilize vibrational overtone
transitions (or low lying electronic transitions) to monitor reagent or product decays.
The trade o is between significantly lower absorption cross sections vs. higher laser
powers and more stable output. A good example can be found in the study on the
HO2 + NO2 reaction by Christensen et al.39 where [HO2] was monitored in the near
IR, either via the OH overtone stretch, or via a rotationally resolved electronic

Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222 | 179


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

transition. The [NO2] was monitored simultaneously by UV absorption. For all IR


absorption techniques a variety of modulation methods can be used to further
improve detection limits.40
An alternative approach to absorption measurements is cavity ring down spectro-
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

scopy, CRDS, first developed by OKeefe41 and with chemical applications recently
reviewed by Marzurenka et al.42 In its simplest form, a pulse of light is injected into
the cavity formed by two highly reflective mirrors. As the light passes back and forth
along the cavity a small fraction leaks out and is detected. The signal on the detector
decays exponentially with a characteristic ring down time, t0. If the cavity now
contains a species which absorbs the light then the ring down time, t, will be lowered
as light is lost by both absorption and cavity loss. The concentration of species, [X],
within the cavity is given by
! "
1 1 L
X# ! 9
t t0 csd
where L is the cavity length, d is the length of the cavity occupied by the absorber, c
is the velocity of light and s the absorption cross section. For kinetic measurements
CRDS can be operated in two fashions; first if the ring down time is short compared
to the timescale of the reaction, then the concentration of the absorbing species will
be essentially constant during the ring down measurement and a pure exponential
decay is obtained. A kinetic trace of reagent or product is obtained by varying the
time delay between the CRDS probe pulse and the photolysis laser. However, if the
timescale of the reaction is comparable with that of the ring down measurement
(usually 110 ms) then the concentration of the monitored species will be changing
during the ring down measurement. In this case Brown et al. have shown, with a
technique that they christened SKaR (Simultaneous Kinetics and Ringdown) that it
is still possible to extract kinetic information for the resulting non-exponential ring
down signals43 as shown in Fig. 4.
A major issue in most absorption measurements including CRDS is how to
determine the absorption pathlength required to calculate absolute concentrations
from a flash photolysis initiated reaction. A variety of dierent geometries have been
tried; a perpendicular arrangement, usually with some shaping of the photolysis
pulse, provides a well determined photolysis region but at the expense of minimising
the overlap between photolysis and probe laser and hence reducing the overall
pathlength. Conversely making the photolysis and probe lasers as co-linear as
possible maximises the overlap (and hence pathlength), but makes it dicult to
accurately determine the actual pathlength.
Optical spectroscopy has many advantages in terms of sensitivity and specificity;
however, it is usually dicult to monitor a range of species using a single piece of
apparatus. In this respect mass spectrometry (MS) has an important advantage as
the technique is essentially universal. A potential drawback is the lack of specificity,
especially with relatively low mass resolution, however this obstacle can generally be
overcome in a variety of ways. The relatively low pressures and non-realtime nature
of discharge flow experiments make the coupling with mass spectrometry relatively
straightforward and examples of branching ratio studies using DF/MS will be
discussed (Section 4.2c). Gutman and Slagle were the main pioneers of LFP/MS
studies using a quadrupole mass spectrometer with photoionization with a resonance
lamp.44 The next major advance came from the coupling of photoionization with
time-of-flight mass spectrometry by Fockenberg et al.45 giving a significant multiplex
advantage, so that all species can potentially be monitoring simultaneously. The

180 | Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222


This journal is '
c The Royal Society of Chemistry 2007
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18. View Article Online

Fig. 4 (a) Ring-down temporal profiles in the absence (upper trace) and presence (lower trace)
of NO3 radicals whose concentration is changing due to reaction. (b) Ratio of the lower to the
upper trace from plot (a) (points) along with an exponential fit (solid line). The discontinuity
that occurs at 12 ms (arrow) results from photolytic production of NO3 from N2O5, and the
somewhat slower return to a horizontal line is due to reactive NO3 loss. Figures reproduced
with permission from ref. 43.

latest development has been to use the Advanced Light Source at Berkeley which
provides tuneable cw synchrotron radiation as the photoionization source for a time-
of-flight mass spectrometer. In general the relatively low energies of photoionization
prevent fragmentation allowing observation of the parent ion of radical species. The
tuneability of the ionization source allows discrimination of species with the same
mass, potentially including dierent isomers.46,47 Synchrotron radiation is not
readily accessible and therefore Blitz et al.48 have developed an apparatus based
on the design by Fockenberg but with photoionization via tuneable VUV radiation
generated by tripling the output of a NdYAG pumped dye laser in a rare gas
medium. As the technique is pulsed, as opposed to the cw radiation of the earlier
design, the duty cycle is lower, but this is compensated by the increased signal to
noise and the flexibility of the tuneable photoionization source.

Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222 | 181


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

2.3 Theoretical calculations

The status of theoretical studies of bimolecular reactions has recently been compre-
hensively reviewed by Fernadez-Ramos et al.49 so only a very brief review will be
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

presented here. The major recent development has been the increasing accessibility
of a variety of relatively high level calculations to experimentalists. Whereas 20 years
ago experimental and theoretical papers on a reaction might be separated by several
years, now it is common place to see theoretical and experimental work in the same
paper giving the reader much greater insight into the mechanism of the reaction and
providing a theoretical basis for extrapolation of the data outside of the experi-
mental range of the study. The synergies between experiment and theory in gas phase
reactions are to be explored in special issue of Physical Chemistry Chemical Physics
edited by Seakins and Robertson to appear in the summer of 2007.
Calculations can broadly be divided into three categories; firstly the calculation of
the potential energy surface, secondly the use of transition state theory and
derivatives to calculate rate coecients and finally calculations associated with
energy transfer between molecules. The latter are particularly important in deter-
mining the product branching ratios of reactions occurring via the formation of an
intermediate complex which may be stabilized (Section 4).
The GAUSSIAN programmes developed by Pople and co-workers have been
heavily used with basis sets of varying sophistication (e.g. G3X50). A computation-
ally cheaper alternative to such wave function theories (WFT) are density functional
theories (DFT), particularly useful for calculating geometries at the saddle points of
reactions. Subsequently more advanced versions of WFT such as multi coecient
correlation methods51 (MCCM) can be used to calculate the energetics of the saddle
points with greater accuracy.
Transition state calculations can be used to predict the rate coecients of
reactions if the partition functions of the reagents and transition state can be
calculated. This is relatively straight forward for reactions proceeding via a well
defined barrier. The increasing accuracy of ab initio calculations means that
predictions are becoming much better, but sometimes the transition state needs to
be tuned to match the experimental data in order to provide predictions of the rate
coecients outside the experimental limits. At low temperatures the role of tunnel-
ling needs to be given greater consideration.52
Many reactions forming complexes proceed without a barrier and hence a well
defined transition state. Under these conditions, the transition state needs to be
located at the minimum flux of reagents and this is found in a variational process.53
Once a complex has been formed, the energy may be considered as being
statistically distributed as in RRKM theory, although examples are known, espe-
cially in small molecules (e.g. HO2,54,55 and HCO56) where there is inecient
coupling between vibrational modes. A competition exists between stabilization of
the complex and dissociation to either reagents or products. Stabilization rarely
involves just a single collision, rather the collisions with a bath gas molecule move
the complex up or down in energy in a stepwise fashion. In order to facilitate
calculations the energy of the complex is divided up into grains. The time evolution
of population in each grain is determined by the master equation which considers
collisional production and removal of the complex from a particular grain as well as
chemical removal via a unimolecular reaction that has an energy dependent (micro
canonical) rate coecient, k(E).5759 An area of current research is how the average
amount of energy transferred in a collision varies as a function of bath gas, complex

182 | Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

and temperature.60 It is now possible to consider reactions occurring through a


variety of complexes and the example of the NH2 + NO reaction is considered in
Section 4.2f.
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

3. Site specific reactions


3.1 Radical + alkane abstraction reactions
Such reactions are important propagation reactions in both atmospheric and
combustion chemistry. In both cases interest centres around the subsequent chem-
istry of the primary, secondary or tertiary radicals that are formed. Under atmo-
spheric conditions with relatively high [NO], primary and secondary radicals are
converted to aldehydes and ketones, respectively as the first step in their complete
oxidation and removal from the atmosphere as shown in Scheme 1.
Aldehydes and ketones have very dierent atmospheric lifetimes with respect to
chemical and photolytic removal processes. The much less reactive ketones can be
transported over much longer vertical and horizontal distances; for example ketones
are important radical sources in the upper troposphere.61
Historically such product studies have been carried by analysing the composition
of the stable end products, usually by gas chromatography coupled to either flame
ionization or mass spectrometric detection. For example, in reaction (5) the possible
products from recombinations of n and iso propyl are n-hexane, 2,3-dimethylbutane
and 2-methylpentane and the product branching ratios of the abstraction reaction
can be obtained with knowledge of the rate coecients for the recombination
reactions. This requirement for additional kinetic information increases the potential
for error; clearly a direct measurement would be advantageous.
A recent example of such a study has been carried out using LFP/diode laser
spectroscopy to monitor the production of HCl from the reactions of Cl atoms with
normal and partially deuterated alkanes.33 In these experiments the reaction of Cl
atoms with the fully hydrogenated alkane is used as the calibration relating the final,
measured HCl concentration with the initial Cl atoms (formed from the photolysis
of a suitable precursor) in a 1:1 manner.
k5
Cl C3 H8 !! HCl propyl R5

The normal propane was then replaced with selectively deuterated propane
(CH3CD2CH3) and the experiment repeated with exactly the same initial Cl atom

Scheme 1 Production of first generation carbonyl products following the reaction of Cl atoms
with i-butane.

Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222 | 183


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

concentration.
k5
Cl CH3 CD2 CH3 !! HCl CH3 CD2 CH3 R5da
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

k5
!! DCl CH3 CDCH3 R5db

Unlike the schematic profiles shown in Fig. 1, the real HCl profiles were biexpo-
nential in nature and the HCl signal was fitted to the following function:
P1
HCl#t e!P2t ! e!P3t 10
P2 ! P3
where P1 is the product of the pseudo-first order rate coecient for H atom
abstraction and the initial Cl atom concentration, P2 is the product of the overall
pseudo-first order rate coecient and the initial Cl atom concentration, and P3 is the
diusional loss term for HCl. Typical experimental traces and bimolecular fits are
shown in Fig. 5.
Kinetic information could be extracted from the P2 parameter yielding bimole-
cular rate coecients which, in the case of room temperature reactions of the normal
alkanes, could be compared with literature data or for the low temperature and
partially deuterated measurements provided new kinetic data. The ratio of the P1
parameters for the normal and partially deuterated measurements directly yielded
the branching ratio to n-propyl production, assuming that the rate coecient for
primary abstraction is not aected by deuteration at the secondary carbon.
The results of the studies on propane and isobutene are summarized in Table 1. In
general the agreement between the direct and indirect studies62,63 is good and the
experimental results provide a good test bed for comparing the predictions of
various structure activity relationships (SAR).6467
Both of the temperature dependent studies show an increasing proportion of
primary abstraction with increasing temperature, however, as the overall rate
coecients for the Cl reactions are either temperature independent (Cl + propane)

Fig. 5 HCl absorption profiles from reaction of identical initial concentrations of Cl atoms
with i-butane and selectively deuterated i-butane. The solid line is the fit to eqn (10). Figure
reproduced with permission of PCCP owner societies from ref. 33.

184 | Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

Table 1 Summary of experimental branching ratios for reactions of Cl atoms with propane
and i-butane and comparison with SAR predictions
Propane Iso-butane
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

Method 195 K 298 K 195 K 298 K

Direct33 40 * 3, 60 * 4 48 * 3, 52 * 3 49 * 4, 51 * 4 62 * 5, 38 * 4
End product 37, 63a 45 * 2, 55 * 2 54, 46a 65 * 2, 35 * 2
analysis63,68 47, 53 64, 36
SAR65 39, 61 57, 43
64
38, 62 57, 43
66
41, 59 59, 41
67
44, 56 62, 38
a
Extrapolation of measured temperature dependence above 298 K.

or negative (Cl + iso-butane), then this implies that rate coecients for the
secondary or tertiary abstractions have a relatively strong negative activation
energy.
The comparison with structure activity relationships is variable. The best agree-
ment is obtained with the earlier SAR of Hooshiyar and Niki66 and Aschmann and
Atkinson.67 The more recent SAR, especially that of Qian et al.,64 have lower rates
for the primary abstraction processes based on a number of recent direct studies on
the Cl + ethane reaction which report values in the region of 5 ( 10!11 cm3
molecule!1 s!1 in contrast to earlier measurements that were closer to 6 ( 10!11
cm3 molecule!1 s!1.
This approach to the study of Cl atom branching ratio reactions was developed by
Taatjes who studied the reaction of Cl with ethanol69 and HD.70 For reaction with
ethanol there is potential for dierent mechanisms to be taking place; a number of
studies have shown unusual rotational71 and vibrational69,72 state distributions of
the HCl product in reactions of Cl atoms with oxygenated species suggesting that the
reaction can sample attractive complexes in the exit channels of the reaction, rather
than a direct abstraction.

3.2 Intramolecular kinetic isotope eects


Radical reactions with HD allow for observation of the intramolecular kinetic
isotope eect and are a good test of the ab initio potential energy surface for the
reaction. The direct study of Taatjes70 showed that the reaction favoured HCl
production (67% at room temperature) and that the branching ratio decreased with
temperature approaching 50% at 550 K.
Intramolecular kinetic isotope eects have been studied for a number of radical
(X) + HD reactions, summarised in Table 2. Branching ratios are obtained from the
ratio of DX:HX and such studies are ideal for relative measurements and highly
precise73 measurements have been obtained often using flow tube measurements.
Intramolecular kinetic isotope eects appear to be divided into several regimes. At
low temperatures the reactions are dominated by tunnelling and transfer of the
lighter hydrogen atom dominates, this process is exemplified by the O + HD
reaction at 300 K where over 90% of the reaction forms OH. In some cases the rate
coecient for the X + HD reaction can be predicted accurately by the mean of the
X + H2 and X + D2 reactions, a good example being the OH + HD reaction. This

Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222 | 185


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

Table 2 Summary of kinetic and branching ratio information on reactions X + HD


Branching
Reaction and k300/cm3 fraction to
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

reference molecule!1 s!1 Arrhenius expression HX at 300 K


# $
3 74 !18 a mol!1
O( P) + HD 3.6 ( 10 8:6 ( 10!12 exp 36:7 kJ
RT 0.93
# $
75 !15 mol!1
OH + HD (3.9 * 0.3) ( 10 5 ( 10!12 exp 17:7 kJ
RT 0.74 * 0.06
# $
mol!1
Cl + HD70 (6.4 * 0.6) ( 10!15 1:8 ( 10!11 exp 19:7 kJ
RT 0.67 * 0.09
# $
32 !14 mol!1
CN + HD (2.1 * 0.2) ( 10 8:2 ( 10!12 exp 14:8 kJ
RT 0.65 * 0.13
# !1
$
73 !11 b
F + HD (1.7 * 0.2) ( 10 1:04 ( 10!10 exp 4:6 kJRTmol
0.59 * 0.01
O(1D) + HD76 2.3 ( 10!10 0.43
a
Extrapolation of temperature dependent data recorded between 422 and 472 K by Presser and
Gordon.77 b Persky.78

implies that the isotopic nature of the spectator atom is irrelevant. If the OH
approaches the H end of the molecule the reaction occurs as if it where the OH +
H2; similarly if the OH approaches the D end of the molecule the rate coecient
can be predicted from the OH + D2 kinetics. For other systems, e.g. Cl + HD, the
measured rate coecients are some 3040% dierent from the means of the X + H2
and X + D2 rate coecients.
As the temperature is increased the reactions become less specific, for example
Taatjes reports the branching ratio for HCl formation to decrease from 0.67 at
300 K to 0.50 at 550 K.70 All the abstraction systems show a decreasing specificity
for H transfer with increasing temperature. Robie et al.74 showed that decreasing
specificity with temperature is what is expected from conventional transition state
theory for the O + HD reaction, however, tunnelling must be included to bring the
experimental and calculated ratios into line.
Finally, an inverse kinetic isotope eect is observed for reactions of O(1D) atoms.
These reactions take place via the insertion of the O atom into the HD bond and
subsequent decomposition of the HOD complex. Break up of the complex on
purely statistical grounds should yield approximately equal concentrations of OD
and OH; the observed preference for OD formation confirms that significant
dynamical constraints are in operation. In conjunction with energy partitioning
measurement, these isotope studies where able to dierentiate between various PES
with the surface developed by Schinke and Lester79 providing the best agreement
with a range of experimental results.

3.3 The reaction of OH with bromopropane


An alternative approach to determine channel eciencies is to study the eect of
isotopic substitution on the overall rate of reaction and the technique is exemplified
in a study by Gilles et al.80 on the reaction of OH with 1-bromopropane, although
the methodology has been used previously.81

OH + CH2BrCH2CH3 - H2O + C3H6Br (R 6)

In essence the technique utilizes the fact that the overall rate coecient is the sum of
the site specific rate coecients and then makes two assumptions regarding isotopic
substitution:

186 | Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

(1) The magnitude of the isotope eect, z, is the same at the a, b, and g positions:
k6c CH3 k6b CH2 k6a CH2 Br
z 11
k6c CD3 k6b CD2 k6a CH2 Br
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

(2) Isotopic substitution at one position has no eect on the site specific
abstraction rate at any other position in the molecule.
Given these assumptions the following equations can be set up for the following
reactions:
OH + CH3CH2CH2Br - products; k1 = k6c(CH3) + k6b(CH2) + k6a(CH2Br)

OH + CD3CH2CH2Br - products; k2 = k6c(CH3)/z + k6b(CH2) + k6a(CH2Br)

OH + CD3CH2CD2Br - products; k3 = k6c(CH3)/z + k6b(CH2) + k6a(CH2Br)/z

OH + CH3CD2CD2Br - products; k4 = k6c(CH3) + k6b(CH2)/z + k6a(CH2Br)/z

OH + CD3CD2CD2Br - products; k5 = k6c(CH3)/z + k6b(CH2)/z + k6a(CH2Br)/z


The overall rate coecients were measured as a function of temperature for the
various isotopomers available. Such calculations require measurements of high
precision and accuracy and great care was taken by Gilles et al. to ensure that
random and systematic errors were minimized. The precision of the measurements
were reported to be B5%. The system described above is over-determined (5
equations and 4 unknowns). Given that a majority of the reaction occurs at the a
and b positions, it would be possible to further refine the calculations and negate the
assumption of equal magnitudes to the isotope eect between the a and b positions,
however, the authors report that a substantial improvement in the precision of the
measurements would be required to add more significance to the branching channel
ratios.
Table 3 summarises the results of the measurements. Reaction is dominated by
abstraction from the b position with abstraction at the a site contributing approxi-
mately 30%. Abstraction from the g, CH3, position is a very minor channel at the
lowest temperatures of the study, B230 K but has a stronger positive temperature
than the other two channels and contributes more significantly (B16%) at the
highest temperature of B360 K.

3.4 Addition reactions


Alkenes react rapidly with a variety of radicals with the initial reaction forming an
addition complex. In some cases where the abstraction generates a relatively stable
radical, abstraction may compete with addition in the kind of reaction addressed in
Section 5.5c. An example is the reaction of Cl atoms with propene where formation
of the resonantly stabilised allyl radical can compete with addition:82

Table 3 Summary of temperature dependence of site specific abstraction reactions for


OH + bromopropane
Temperature/K k6c/k6 k6b/k6 k6a/k6

230 0.08 * 0.08 0.61 * 0.02 0.30 * 0.08


300 0.12 * 0.04 0.56 * 0.02 0.32 * 0.04
360 0.16 * 0.02 0.52 * 0.02 0.32 * 0.02

Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222 | 187


This journal is '
c The Royal Society of Chemistry 2007
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18. View Article Online

Scheme 2 First generation products from the reaction of OH with propene.

Cl + C3H6 - C3H6Cl (R 7a)

- HCl + C3H5 (R 7b)

For unsymmetrical alkenes or alkynes the position of the initial addition can
determine the final product distribution. In general the radicals will attack the least
substituted position of the multiple bond, forming the more stable radical, however,
this process may not be specific. The predominant fate of alkenes released into the
atmosphere is reaction with OH. For a simple unsymmetrical alkene such as propene,
the position of the attack has no eect on the main products formed. Following
addition, the adduct radical will rapidly form a peroxy radical and, except in pristine
environments, this will react with NO. Both of the resulting b-hydroxyalkoxy radicals
will decompose to yield formaldehyde and acetaldehyde as shown in Scheme 2.
For the corresponding alkyne, propyne, the situation is slightly more complex.
The peroxy species still contains a multiple bond and can undergo rapid internal
rearrangements prior to reaction with NO. One possible rearrangement yields
methyl glyoxal from either peroxy radical, however the alternative rearrangement,
which involves the formation of a cyclic intermediate, will form either formic or
acetic acid depending on the position of the initial OH addition. The corresponding
co-products (CH3CO and HCO) have dierent ozone generating capabilities.
Hatakeyama et al.83 studied the products of the OH addition to propyne (Scheme
3), in a chamber experiment where stable products where monitored by FTIR. The
dominant product was methylglyoxal (0.53 * 0.03) which can originate from
addition at either site, but of the other products, only formic acid was observed
suggesting that the initial addition occurred entirely at the terminal position.
However, a more recent study by Yeung et al.,84 using a discharge flow system with
mass spectrometric detection, observed both formic and acetic acid. Although the
yields were not quantified, acetic acid was reported as the dominant acid.

Scheme 3 First generation products from the reaction of OH with propyne.

188 | Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

The oxidation of alkynes is a topic of current interest. Alkynes are considered to


be useful tracers for anthropogenic activity or biomass burning,85 the di-carbonyl
species formed during atmospheric oxidation may play important roles in aerosol
formation and growth and in general, the mechanisms of acid formation in the
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

atmosphere are poorly characterized. Clearly there is scope for additional work in
this area.
The final reaction briefly considered in this section is the OH initiated decom-
position of isoprene. Isoprene is one of the major biogenic emissions and is a major
source of formaldehyde, CO and oxygenated species in the atmosphere. Reaction
with OH is the major daytime loss process; the unsymmetrical nature of this diene
means that there are four dierent addition sites. It is beyond the scope of this review
to consider the oxidation pathways in any depth, however, the predicted yields of the
major products methacrolein and methylvinyl ketone are dependent on the initial
position of OH addition.86,87 The dierence in chemical lifetimes of aldehydes and
ketones has been discussed earlier in Section 3.1.

4. Complex forming reactions


4.1 Introduction
This is the largest type of reaction considered in this review. The reactions take place
on a potential energy surface shown schematically in Fig. 6. The details of the PES
may vary, but the major common theme is that a complex is formed with the release
of a significant amount of chemical energy with the formation of new bonds. The
complex formed is therefore highly chemically activated and may be able to access a
variety of dierent product channels. Whether or not all the thermodynamically
allowed products can be accessed depends on the lifetime of the adduct and the
barriers to product formation; given a relatively long lifetime vibrational energy flow
within the adduct will occur and the product distribution is likely to be statistical in
nature, reflecting the relative density of states at the transition states to product
formation. Under certain conditions the lifetime of the adduct may be sucient such

Fig. 6 Schematic potential energy surface for a multichannel reaction proceeding via complex
formation.

Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222 | 189


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

that collisional stabilization can compete with unimolecular dissociation of the


adduct and the stabilised adduct becomes one of the product channels. At the other
extreme, the lifetime of the complex may be very short and the complex is not able to
explore all the available regions of phase space and the product distribution is
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

controlled by dynamical rather than statistical constraints.


If, as shown in Fig. 6, the exit barriers are significantly lower than the entrance
channel then re-dissociation of the complex back to reagents is insignificant
compared to the forward channels. However, in circumstances where the entrance
and exit barrier heights become comparable, redissociation to reagents may become
significant and the kinetics of reagent removal will become pressure dependent.
The mechanism by which the complex is formed may vary quite significantly, for
example the complex may be formed from the barrierless recombination of two
radicals, addition (activated or not) to multiple bonds or insertion of highly reactive
radicals such as CH or 1CH2 into bonds.
Our review of product studies of reactions proceeding via complex formation is
divided into two components. Firstly we consider examples of studies that illustrate
the characteristic features of the PES and then secondly, we consider a number of
topical examples.

4.2 Characteristics of complex forming reactions


4.2a Forming the complex. The recombination of two doublet species to form a
singlet complex in a barrierless reaction is the simplest example of complex
formation and is exemplified by the recombination of alkyl radicals.88,89 For methyl
radical recombination there are no open forward channels and stabilization of the
ethane complex is the sole mechanism for methyl radical removal. Another example,
considered in more detail below, is the reaction of CH3 radicals with oxygen atoms
forming a methoxy complex. The recombination of halogen oxides, both self and
cross reactions, have received significant study due to their role in stratospheric
ozone depletion and the area has been reviewed by Bedjanian and Poulet.90 The self
reaction of ClO is the prototypical reaction; however, in a slight complication to the
above reactions, it is possible to form two dierent initial complexes ClOOCl and
ClOClO in barrierless reactions as shown in Fig. 7, a PES calculated by Zhu and
Lin.91 Theoretical calculations91,92 suggest that the former is the dominant process
and the calculated product yields are in reasonable agreement with experiment.93
The addition of radicals to a double bond is another way in which a complex can
be formed. This can either be an activated or non activated process, an example of
the latter is the addition of CN radicals to alkenes where the kinetics are pressure
independent and show a negative temperature dependence characteristic of barrier-
less reactions.94,95 As discussed above, for unsymmetrical multiple bonds the site of
the addition will determine the product spectrum. For some addition reactions there
is the possibility of the reaction being influenced by van der Waals complexes in the
entrance channel.96,97
Other complex forming reactions are more intricate in nature, but despite the
significant amount of bond breaking and forming processes, also proceed without a
barrier. Good examples can be found in the reactions of CH radicals, an important
intermediate in combustion and in the atmospheres of some of the outer planets. The
reaction of CH with methane has received significant attention as a relatively simple
reaction and also as a potential calibrant for other CH reactions.98100
CH + CH4 - C2H5* - H + C2H4 (R 8)

190 | Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222


This journal is '
c The Royal Society of Chemistry 2007
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18. View Article Online

Fig. 7 Schematic of the singlet PES for the ClO self reaction computed at the G2M level.
Figure reproduced with permission from ref. 91.

The reaction is thought to proceed via insertion of the CH radical into a CH bond
in the methane with evidence coming from a relatively large isotope eect. As CH4 is
replaced by CD4 the observed rate coecient decreases by approximately 33%.98,99
The reaction has been modelled by Taatjes and Klippenstein101 who find good
agreement between experiment and theory for the CH4/CD4 ratio, but with an
inverted prediction for the eect of replacing CH with CD. They argue that this
incorrect prediction of secondary isotope eects indicates the need to consider the
variational nature of such transition states in more detail.

4.2b Thermal decomposition of AB. It is also possible to access the products


shown schematically in Fig. 6 by thermal activation of the AB species. The product
branching ratios of thermal decomposition reactions should possibly be considered
in a completely separate section, however, such reactions are closely related to the
decomposition of chemical activated complexes.
In some cases it is possible to compare the product distribution of chemical
activated molecules with thermal distributions. The insertion of the CH radical into
the CH bond of ethane generates a chemical activated n-propyl, n-C3H7*, radical
with approximately 400 kJ mol!1 of internal energy.
CH + C2H6 - C3H7* - products (R 9)
The n-propyl radical can decompose via two channels or isomerize to the slightly
more stable i-propyl radical:
n-C3H7 - CH3 + C2H4; Ea = 122 kJ mol!1 (R 9a)
!1
- H + C3H6; Ea = 143 kJ mol (R 9b)
!1
- i-C3H7; Ea = 161 kJ mol (R 9c)
The H atom yield from reaction (9) has been measured by two groups and values of
14 * 6%102 and 22 * 8%103 have been obtained using complimentary techniques
(laser flash photolysis and flowtube methodology). In comparison, a direct study of
n-propyl decomposition via shock tube measurements yields less than 5% H at the
highest temperature (1500 K) of the study.104 The two studies of reaction (9) are

Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222 | 191


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

consistent within errors, but the interpretations of the results are slightly dierent.
McKee et al.102 calculate product branching ratios from microcanonical rate
coecients obtained from an inverse Laplace transformation of Arrhenius repre-
sentations of reactions (9ac) from Yamauchi et al.104 The calculated branching
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

ratio was 23% and varying the relative activation energies for the dissociations by
*4 kJ mol!1 gave branching ratios between 17 and 31% which McKee et al.
considered to be consistent with the experimental observations. Thus McKee et al.
consider their results to be compatible with those obtained from the thermal
decomposition; chemical activation being equivalent to very high temperature
dissociation. Galland et al.103 carried out a RRKM calculation based on an ab
initio surface which they calculated and determined a yield of 12% from decom-
position of the n-propyl radical. They considered that this result was incompatible
with their observations arguing that approximately 10% of the CH + C2H6 reaction
proceeds via insertion into the more sterically hindered CC bond forming the
i-propyl radical which decomposes solely to H + C3H6 and therefore only 4% of the
observed yield arises from reaction (9b). Both calculations agree that isomerization
of the n-propyl radical cannot compete with direct dissociation.

4.2c Relative heights of entrance and exit barriers. If the barriers of the entrance
and exit channels are comparable in height, then unimolecular dissociation of the
complex back to reagents may compete with either stabilization or forward
dissociation to products. One example of this are the RCO + O2 reactions
considered below in Section 4.2d, where the kinetics are pressure dependent.
Another good example is the isotopic substitution of deuterium into the methyl
radical via the CH3D* intermediate:

Clearly the barrier heights for the entrance and exit channels are equal, however,
dierences in zero point energy ensure that the forward reaction with H elimination
is favoured. The CH3 + H reaction is an important testbed for theories of
recombination reactions,105 however, the small size of the system means that the
high pressure limiting rate coecient (kN) is only realised at very high pressures,
with the radical radical nature of the reaction providing further experimental
complications. Seakins et al.106 studied the isotopic reaction as an alternative
method of obtaining kN, building on earlier work by Brouard et al.107 The reaction
was carried out in a discharge flow system with the reagents being generated by F
atom (from a microwave discharge) reactions at the tip of a sliding injector.
F + D2 - DF + D (R 11)
F + CH4 - HF + CH3 (R 12)
D atoms were generated in large excess and hence the decay of the methyl radical, as
monitored by mass spectroscopy, took place under pseudo-first order conditions.
Monitoring mass numbers corresponding to CH3, CH2D, CHD2 and CD3 showed
the gradual sequential replacement of H with D as shown in Fig. 8.
Master equation calculations were carried out to determine the fraction of the
methane complex that redissociates back to reagents for each of the methyl
isotopomers, with this information it was possible to correct the rate coecients
for each of isotopomers to give an averaged value for k1 CH3 H of (2.9 * 0.7) (

192 | Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222


This journal is '
c The Royal Society of Chemistry 2007
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18. View Article Online

Fig. 8 Plot showing the gradual conversion of CH3 to CD3 via sequential replacement of H by
D. The growth and decay of the intermediate CH2D and CHD2 radicals can clearly be seen.
Adapted from ref. 106.

10!10 cm3 molecule!1 s!1 in good agreement with theory and importantly the
108
predicted positive temperature dependence for k1 CH3 H is confirmed.
The reaction of CH with H2 is another reaction with similar entrance and exit
barriers. As discussed earlier, CH reactions often proceed via an insertion and this
reaction appears to be no exception with the resultant CH3, formed in a barrierless
reaction, either being stabilised or dissociating to give CH2 + H in a slightly
endothermic reaction. Brownsword et al.109 studied the reaction over a wide range of
temperatures and pressures. With such a small system very high pressures are again
required to determine k1 CHH2 , Brownsword et al. used similar isotopic studies to
those described above (CH + D2) to obtain the high pressure limit, but also used the
reaction of vibrationally excited CH to probe k1 CHH2 . The use of vibrationally
excited radicals to probe the high pressure limit of barrierless reactions has been used
by Smith and co-workers on a number of occasions.110 The assumption is that the
presence of vibrational energy in the spectator bond of the radical does not influence
the formation of the complex, however, when formed, the complex is statistically
much more likely to produce ground state species rather than to redissociate to the
vibrationally excited radical. Therefore the loss of CH(v) should be equate to the rate
coecient of formation of the complex, k1 CHH2 . Brownsword et al. were able to
place the CH + D2 and CH(v) + H2/D2 reactions on a common scale via transition
state theory and the resultant estimates of k1 CHH2 are in reasonable agreement and
show a negative temperature dependence consistent with barrierless formation of the
methyl complex.
Studies of the CH + H2 reaction as a function of pressure produced slightly more
controversial results. At high temperatures and low pressures the rate coecients
showed a positive temperature dependence, consistent with the endothermic forma-
tion of CH2 + H. At low temperatures where the endothermic channel is inacces-
sible, the rate coecient showed a conventional increase with pressure (4400 Torr).
However, at high temperatures (4500 K), where the endothermic product channel is

Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222 | 193


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

more accessible, the rate coecients showed an abnormal variation with pressure,
firstly decreasing over the range (4B100 Torr) and then increasing at higher
pressures.
Brownsword et al. interpreted this negative temperature dependence in terms of
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

two dierent methyl radical species, CHaa 3 , chemically activated with sucient
energy to dissociate to CH2 + H and CHa 3 , chemically activated, but below the
barrier to form CH2 + H. The authors then argued that it is possible that the overall
rate is reduced more by partial collisional deactivation of CHaa 3 to CHa 3 (from
which redissociation is still possible) than it is increased by stabilization of CHa 3 to
CH3, although full master equation calculations are really required to verify this
hypothesis.

4.2d Complex stabilization vs. dissociation. If energy removal can compete with
either forward or reverse dissociation of the complex, then the complex may be
stabilised and become one of the products. A good example of this can be seen in the
addition of CN radicals to a variety of alkenes and acetylene.111 For the addition to
acetylene and ethene, the complex is not stabilised at pressures below 200 Torr and
100% of the reaction goes forward with the elimination of an H atom, however, as
the size of the complex increases (e.g. addition of CN to propene and iso-butene) the
yield of the forward reaction to produce H decreases as a function of pressure as
shown in Fig. 9. The rate coecient for the loss of CN is pressure independent; the
competition is now solely between stabilization and forward dissociation.
Another class of such reactions that is currently receiving much investigation is the
addition of oxygen to a variety of radicals. Taatjes has recently reviewed the
reactions of alkyl radicals with oxygen112 and therefore such systems will not be
considered in this work. However, complex reaction mechanisms for the addition of
oxygen to a variety of radicals are being increasingly observed in laboratory studies.
A good example is the reaction of the acetyl radical, CH3CO, with oxygen. The first
evidence that this reaction might be anything other than a simple termolecular
recombination came from work by Michael et al.113 who showed that at low pressure
and in the presence of oxygen there was eectively no loss of the OH radical from

Fig. 9 Pressure dependence of the H atom yields for reactions CN + C3H6 () and CN +
iso-C4H8 (&) at 298 K.

194 | Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

reaction with excess acetaldehyde. Michael et al. interpreted this observation as


being due to OH regeneration from the acetyl + O2 reaction.
OH + CH3CHO - H2O + CH3CO (R 12)
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

CH3CO + O2 - OH + products (R 13)


114
OH production was confirmed in a chamber study by Tyndall et al. The results
were interpreted via the following mechanism:

At atmospheric pressure there was essentially 100% production of the acetylper-


oxy radical (R13b) and so reaction (13a) is not an important tropospheric source of
OH. Acetylperoxy yields decreased with pressure to about 50% at the lowest
pressure studied (B6 Torr). Interestingly the co-product of the reaction could not
be identified.
Subsequently a number of other measurements have been made on the system.115
Although not of direct atmospheric importance, the reaction is an excellent marker
for the acetyl radical,116,117 which is otherwise dicult to detect. The co-product of
the OH channel has still to be positively identified.118

4.2e Statistical or dynamical control over product distributions. Studies on the


CH3 + O system have helped to develop some new ideas on the way that reactions
can take place, specifically that reactions need not always proceed through the
transition state for the reaction but rather over a ridge.119
Early experimental studies showed that CH3 + O is a fast reaction (k = 1.3 (
10!10 cm3 molecule!1 s!1) that is virtually temperature independent.120 The fast
kinetics were attributed to the barrierless formation of a methoxy, CH3O, inter-
mediate which dissociates to form formaldehyde and H. The formaldehyde product
was observed via both mass spectroscopy120 and LIF.121 However, in 1992 Seakins
and Leone122 reported the observation of a new channel in the reaction, the direct
production of vibrationally excited CO via time resolved FTIR with H2 and H as the
implicit co-products.
CH3 + O - HCHO + H; DrH = !286 kJ mol!1 (R 14a)
!1
- HCO + H2; DrH = !351 kJ mol (R 14b)
!1
- H + CO + H2; DrH = !287 kJ mol (R 14c)
- OH + CH2; DrH = +34.5 kJ mol!1 (R 14d)
By comparing the CO signal with that produced from acetone photolysis, Seakins
and Leone estimated that the branching ratio to channel (14c) was 0.40 * 0.10.
Subsequent, more precise measurements by Fockenberg and co-workers using
TOFMS123 and diode laser absorption124 have confirmed the production of CO,
but suggest that the yield is closer to 0.2. More recently, Leone and co-workers125
revisited the reaction probing the eects of deuteration of the methyl radical on the
vibrational distribution of the CO product, in conjunction with high level theoretical
calculations, to investigate the dynamics of the reaction.
CO was presumed to be formed in a two step process involving the initial
elimination of H2, followed by the rapid dissociation of the HCO radical (the

Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222 | 195


This journal is '
c The Royal Society of Chemistry 2007
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18. View Article Online

Fig. 10 B3LYP/6-31G* energy contours for CH3O as a function of two of the CH bond
lengths. All other geometric parameters are optimized at each point. The darkest contours
correspond to the energy of the saddle point for H + H2CO - H2 + HCO. The lightest
contours are higher in energy and the darker contours are lower. The contour increment is
5 kcal mol!1. The length scales are in A. Figure reproduced with permission from ref. 125.

CH bond strength in HCO being very weak), however, no transition state could be
located for the initial H2 elimination. The calculations of Marcy et al.125 confirmed
the absence of a direct transition state (or saddle point) linking methoxy and HCO
+ H2. A transition state was located for H2 elimination from the CH2OH isomer,
however, the barrier for this process was high compared to direct H elimination from
either CH3O isomer and would not be expected to compete.
The explanation for the observation of the CO product came from dynamics
calculations run on the high level PES calculated by Marcy et al. Fig. 10 shows a
schematic of the PES for product formation from the methoxy radical.
The methoxy radical is the well in the bottom left hand of Fig. 10 and the valleys
parallel to the axis represent two pathways for the loss of H to H + HCHO. The
HCO + H2 channel is the broad valley leading to the upper right hand of the figure
and is separated from the methoxy well by a ridge. Dynamics calculations show
that this valley can be accessed indirectly by trajectories that leave the methoxy well
and initially proceed towards H + HCHO production, but then cross a saddle
point representing H atom abstraction from HCHO. The results of a number of
dynamics simulations suggest that the HCO + H2 yield is 0.14 * 0.03 in excellent
agreement with the measurements of Fockenberg et al. Trajectories for the CD3
system showed a decrease to 0.09 * 0.02 in quantitative agreement with the
experimental measurements.
The product distribution of the CH3 + O reaction therefore appears to be
controlled by the dynamics of the process rather than by a statistical partitioning

196 | Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

of energy. It has been postulated that such processes may be relatively common.
Townsend et al.119 have carried out experimental and dynamical studies on
formaldehyde dissociation near the threshold for H + HCO production. Excited,
ground electronic state HCHO was generated from laser excitation (as opposed to
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

chemical activation in the systems discussed in this review) and the H2 and CO
product distributions were observed by ion-imaging techniques which allow for
quantum state characterization of the products. A bimodal distribution was
observed; highly rotationally excited CO correlated with low vibrational excitation
of H2 in one mode, in the other mode the CO was rotationally much less excited, but
the H2 was excited up to V = 7. Dynamical calculations show that as for H2 + HCO
production from CH3O, the reaction starts with dissociation towards H + HCO,
but that the H atom samples a wide range of reaction space before internally
abstracting the H from the incipient HCO fragment. The process occurs at extended
HH distances leading to the highly vibrationally excited H2 observed.

4.2f Multiwell reactions. In many cases there are several intermediates between
reagents and products and modelling such systems can prove particularly challen-
ging. In the master equation formalism this requires the consideration of micro-
canonical rate coecients into the accessible wells (including reverse reactions) in
addition to collisions stabilization/excitation and production of products/reagents if
accessible from that well. Details on the master equation modelling of such reactions
can be found in a review by Pilling and Robertson.57
An important example is the reaction of NH2 and NO which is a crucial reaction
in the thermal DeNOx process and the reaction coordinate diagram126 is shown in
Fig. 11.

Fig. 11 Reaction coordinate diagram for the NH2 + NO reaction. Figure reproduced with
permission from ref. 126.

Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222 | 197


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

The crucial channels are:

NH2 + NO - N2H + OH (R 15a)


Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

- N2 + H2O (R 15b)

Reaction (15a) is a chain branching process as the N2H radical formed quickly
decomposing to give H atoms, whereas reaction (15b) is chain terminating with the
formation of stable unreactive species.
The reaction has been studied by a number of groups, but the most recent work is
an elegant study by Song et al.23 who did not measure the branching ratio by direct
observation of the products, but rather by modelling the profile of NH2 in the
presence of excess NO and NH3. The radical products from reaction (15a) react
rapidly with ammonia, regenerating NH2. Reaction conditions are chosen carefully
so that NH2 profile is most sensitive to the branching ratio, a, for the two channels as
shown earlier in Fig. 2b.
Fig. 12, from ref. 23 summarizes the previous experimental work and comparison
with theoretical calculations. Both theory and experiment now appear in excellent
agreement and reviews on the theoretical aspects of the reaction can be found in
work by Miller and colleagues.49,127

4.3 Specific examples


4.3a HCCO + O2. This reaction is an example of a complex forming system
where the dominant products (H + CO + CO2) can be formed via two dierent

Fig. 12 Summary of temperature dependent branching ratio data, a, for the NH2 + NO
reaction from ref. 23. Information on references can be found in ref. 23. Figure reproduced with
permission from ref. 23.

198 | Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

pathways out of the initial complex. Practically, the reaction is of importance in


explaining the prompt formation of CO2 in the oxidation of acetylene. The original
explanation of the phenomena was based on the production of methylene from the
O + C2H2 reaction (R16), however experimental128 and theoretical studies128,129
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

have shown that the production of HCCO + H is the dominant channel despite
being the less thermodynamically stable channel:
O + C2H2 - HCCO + H; DrH = !82 kJ mol!1 (R 16a)
!1
- CH2 + CO; DrH = !200 kJ mol (R 16b)
130
Klippenstein et al. then proposed that the reaction of HCCO with oxygen (R17)
could be the source of CO2. The reaction has a number of exothermic channels:
HCCO + O - H + CO + CO2; DrH = !462 kJ mol!1 (R 17a)
!1
- OCHCO + O; DrH = !5.4 kJ mol (R 17b)
!1
- OH + 2CO; DrH = !360 kJ mol (R 17c)
Klippenstein et al. suggested that reaction (17a) was the dominant channel (480%),
but showed that the potential energy surface for the reaction was complex with two
possible channels to the dominant products (Fig. 13).
The relative importance of the three and four membered transition states to the
products has been investigated by Zou and Osborn131 via isotopic substitution of the
molecular oxygen. As shown in Fig. 13, the products from the two dierent
mechanisms are isotopically distinguishable. Zou and Osborn used high resolution
time resolved Fourier Transform IR (TRFTIR) emission spectroscopy to monitor
the evolution of the products following the photolytic generation of the HCCO

Fig. 13 Potential energy diagram including zero point energy for the HCCO + 18O2 reaction
taken from refs. 130 and 131.

Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222 | 199


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

(from ethyl ethynlether) in the presence of either 16O2 or 18O2. High resolution
experiments (0.115 cm!1) were required to resolve individual rovibrational lines; the
CO spectra were quite complex as there was emission from a large number of
vibrational levels and the experiments were challenging. Zou and Osborn were able
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

to definitively assign transitions associated with C16O and C18O as shown in Fig. 14
and 18OC18O and 16OC18O and hence measure a branching ratio channel (17a) via
the 3 and 4 membered transition states with the latter accounting for over 85% of the
reactive flux.
TRFTIR emission spectroscopy is a powerful tool for such systems132 as it is also
possible to determine the nascent vibrational distribution of the CO and CO2
products. The observed vibrational distributions can be fitted with a vibrational
temperature (CO, 9000 * 1500 K, CO2, 6100 * 700 K) consistent with the
statistical distribution of energy between the products.

Fig. 14 Emission spectrum of C18O from HCCO + 18O2 with the corresponding fit and
residuals. Figure reproduced with permission of PCCP owner societies from ref. 131.

200 | Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

4.3b C(3P) insertion reactions. The insertion of reactive radical species (C, CH,
1
CH2) into saturated and unsaturated hydrocarbons are good examples of addition/
elimination systems, many of which have complex PES with several possible
intermediates. A number of systems have received extensive coverage due to the
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

importance of the reactions in interstellar chemistry, planetary atmospheres and


combustion (particularly in soot formation).
The reaction of C(3PJ) with allene, C3H4, is typical of such reactions and has been
studied by a number of complimentary techniques. C(3PJ) has been observed in high
abundance in dense interstellar clouds and reactions with unsaturated species of the
type:

C(3PJ) + CnHm - Cn+1Hm!1 + H (R 18)

have been proposed as routes to the synthesis of larger carbon based species.133
Early room temperature measurements by Husain and co-workers134 showed that
the reaction was fast and low temperature Laval experiments by Chastaing et al.135
showed that the reaction remained fast over the temperature range 15295 K ((3.5 *
0.8) ( 10!10 cm3 molecule!1 s!1) demonstrating that there is no barrier to complex
formation. In a complimentary study Chastaing et al. used a cross molecular beam
apparatus to explore how the reactive cross section for H atom production varied
with the relative translational energy. The resulting cross sections varied as e!0.55
tr ,
which when one takes account of the T0.5 dependence of the collision rate yields an
essentially temperature independent rate coecient in agreement with the kinetic
studies.
There are three possible exothermic channels for the reaction:136

C(3PJ) + CH2QCQCH2 - n-C4H3 + H; DrH = !180 kJ mol!1 (R 19a)

- HCCCCH + H2; DrH = !450 kJ mol!1 (R 19b)


!1
- HCCCCH + 2H; DrH = !18 kJ mol (R 19c)

Kaiser and co-workers136 also used a cross molecular beam apparatus to study the
reaction, but with mass spectrometric detection. Peaks were observed at m/z = 51
(C4H3) and also at mass 50 suggesting the possibility that channels (19b) and (19c)
may compete. However, the behaviour of the peak at mass 50 is identical to that at
51 suggesting that it comes from fragmentation in the ionization region and that
channel (19a) is the sole product, however, Chastaing et al. point out that
sensitivities to species (e.g. C4H2 + H2) that will be scattered over a wider range
of laboratory angles will be lower. The scattering angles are consistent with the
formation of a long-lived C4H4 complex with a lifetime comparable or greater than
the rotational period.
Finally, Loison and Bergeat137 have examined the reaction in a fast flow system at
2 Torr with LIF detection of the C(3PJ) removal and H atom production. The time
resolved study of both reagent and product ensures that the observed product does
originate from the primary reaction. H atom signals are placed on an absolute scale
by the use of a calibration reaction, in this case with ethene (96% H production). For
reaction (19), the H atom yield is essentially 100% and in combination with the
previous studies discussed above, it is clear that reaction (19a) is the sole active
forward channel in the reaction. Not surprisingly given the low pressures of the fast
flow system, stabilisation is not significant.
Fig. 15 shows the PES for the reaction calculated by Mebel et al.138 The ab initio
study confirms the experimental result in that C4H3 + H are predicted to be the only
products with over 98% of the C4H3 being produced in the most stable n-C4H3

Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222 | 201


This journal is '
c The Royal Society of Chemistry 2007
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18. View Article Online

Fig. 15 Profile of the PES for the C(3PJ) + allene reaction calculated at the G2M(RCC,MP2)
level. Figure reproduced with permission from ref. 138.

isomer (p1 in Fig. 15), predominantly formed directly from c1. p2 is the only other
isomer formed with any significant yield.

4.3c Ozone alkene reactions. These reactions are important loss processes for
anthropogenic and biogenic alkenes in the atmosphere. More recently the reactions
have received further investigation as the systems have been shown to be important
HOx (OH and HO2) radical sources (e.g., ref. 139142) and unlike a majority of
atmospheric HOx formation processes the ozone alkene source will remain active at
night time. The reaction has been suggested as one of the major explanations for the
observation of high night time HOx in the PROPHET campaign which took place in
a forested environment with high concentrations of biogenic alkenes.143 The
products of the reaction are also associated with aerosol formation144 with con-
sequent issues for both air quality and climate change. Ozone alkene reactions are
probably at the end of the spectrum of simple gas phase reactions, but their
importance warrants consideration in such a review. An enormous amount of work
has been carried out and only the briefest outline can be presented in this review. The
interested reader is referred to a very recent review by Marston and Johnson.145
OH generation is thought to take place in a two stage mechanism with the first
step being the production of a Criegee intermediate (CI) and a stable carbonyl
(Scheme 4). For unsymmetrical alkenes two dierent carbonyls can be formed. The
vibrationally excited CI can directly unimolecularly decompose or can be collision-
ally stabilized with subsequent unimolecular or bimolecular reactions. The

Scheme 4 Ozonolysis mechanismCriegee formation.

202 | Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

competition between stabilization and decomposition leads to a pressure dependence


in the observed OH yields.142 OH formation takes place via a cyclic transition state
to an excited hydroperoxide which rapidly decomposes to OH (with the co-product
leading to highly oxygenated species). Alternatively the CI can undergo ring
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

formation to form a dioxirane.


The OH yield from ozone alkene reactions has been studied by a number of
methods, both direct and indirect. Many of the studies have involved end product
analysis of stable species, as opposed to the direct, time resolved measurements
mainly discussed to date. For such large systems, time resolved spectroscopic
techniques are not always applicable and more information may be obtained from
chromatography or FTIR analysis of stable products. Direct radical detection and
end product analysis of stable products are often complementary in the information
that they generate and a number of chambers capable of making both types of
measurements have recently become available.
Indirect studies on the OH yields have involved tracer type measurements. An
excess of hydrocarbon(s) reactive towards OH, but not to ozone, is added to the
mixture. The OH yield can be identified by either measuring the increase of a known
product (e.g. cyclohexanone from OH attack on cyclohexane146) or from the
depletion of a hydrocarbon such as trimethylbenzene (TMB).140 Neither technique
definitively identifies OH; other radicals could be initiating the hydrocarbon oxida-
tion. Paulson and co-workers introduced a mixture of hydrocarbon species, looking
at the relative loss of species which will be characteristic of the attacking radical
providing more definitive proof OH production.147 Table 4 (adapted from ref. 140)
illustrates the range of uncertainties from such reactions and more studies are
required to decrease the uncertainties.
The results from Gutbrod et al. are consistently lower than other determinations,
possibly the conversion of CO to CO2 is a less specific probe of the OH yield.
Otherwise, for the simple alkenes, the results from a number of groups are in
excellent agreement; however, for the larger, biogenic species (which account for a
majority of the global alkene emissions) there are still discrepancies.
Direct observation of OH production has been pioneered by the Harvard group
using a high pressure flow tube system with observation of the OH product by LIF
e.g.142,152,153 OH radicals are observed in the steady state:
O3 + alkene - a[OH] + products (kO3) (R 20)
OH + alkene - products (kOH) (R 21)

Table 4 OH yields from a series of ozone alkene reactions at atmospheric pressure

Alkene Rickarda 140


Atkinsonb 143
Paulsonc149 Moortgatd 150
Gutbrode 151

Ethene 0.14 0.12 0.18 0.08


Propene 0.32 0.33 0.35 0.34 0.18
E-2-butene 0.54 0.64 0.24
Isoprene 0.44 0.27 0.25 0.26 0.19
b-pinene 0.24 0.35
a-pinene 0.83 0.85 0.70
a
TMB removal. b Production of cyclohexanone. c Relative rate removal of hydrocarbons.
d
Change in the observed kinetics of alkene removal. e Conversion of CO to CO2.

Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222 | 203


This journal is '
c The Royal Society of Chemistry 2007
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18. View Article Online

Fig. 16 Typical run, reproduced with permission from ref. 142, for ozone + tetramethy-
lethene at 10 Torr. Ozone is modulated at five dierent concentrations, with the resultant OH
concentration varying accordingly. OH yield is the slope of the [OH] vs. [O3] line, scaled by
kOH/kO3.

so that:

akO3 O3 #
OH#ss 12
kOH

Fig. 16 shows the precision of a typical experiment and the resultant plot of eqn (12)
to give the yield. The yields are strongly pressure dependent and can be reproduced
relatively well via a master equation calculation.
Scheme 5 shows that OH is generated from a cyclic transition state; such a
transition state will be highly strained when the R group is an H, e.g. for O3 +
ethene, however, Table 4 shows a consistent and significant OH yield for this
reaction. Kroll et al.152 have used selective deuteration to probe possible mechanisms
for such reactions (Scheme 6).

Scheme 5 Decomposition of the Criegee intermediate.

Scheme 6 The use of dierent isomers to probe reaction mechanisms.

204 | Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222


This journal is '
c The Royal Society of Chemistry 2007
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18. View Article Online

Scheme 7 Possible routes to OH formation. From ref. 152.

OH and OD yields are measured in a similar fashion to that described already and
Kroll et al. suggest that there is a minor channel to OH formation from the dioxirane
(Scheme 7) and neither syn-anti isomerisations (Scheme 6), four membered transition
states or direct decomposition of the ozonide (Scheme 7) contribute to OH
production in such systems.

4.3d The reaction of H (and CH3) with O2. It seems appropriate to conclude our
discussion on complex forming reactions with a system, that of H + O2, that formed
one of our initial introductory examples:
k2a
H O2 !! OH O R2a

k2b
!! HO2 R2b
The reaction is of enormous importance in combustion. Both reactions proceed on
the same potential energy surface with a deep minimum corresponding to HO2
formation. The initially formed complex can either be stabilized, redissociate or
proceed to the endothermic OH + O asymptote. Reaction (2a) is a chain branching
step two radicals are produced for every H atom consumed increasing the number of
chain carriers; conversely reaction (2b) is essentially chain terminating due to the
relatively low reactivity of the HO2 radical and its ability to take part in recombina-
tion reactions such as
HO2 + HO2 - H2O2 + O2 (R 22)
H + HO2 - H2 + O2 (R 23)
At 1100 K, Maas and Warnatz154 showed that reactions (2a) and (2b) have
approximately equal, but opposite sensitivities to determining ignition delay times
of a H2/O2 system.
Despite the apparent simplicity of the reactions, the H + O2 system poses
significant experimental and theoretical challenges. The recombination reaction is
far from the high pressure limit and deactivation by various bath gases needs to be
considered explicitly155,156 over a wide range of temperatures. Improvements in the
measurement techniques associated with shock tube measurements e.g., ref. 157 and
in high pressure flash photolysis measurements9,158 mean that models are highly
constrained and details of the mechanism can be investigated. Reaction (2a) now

Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222 | 205


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

appears to be well characterised over a very wide temperature range (9005000 K),
but there is still interest in the non statistical behaviour of the small HO2* complex
e.g., ref. 54 The current status has recently been reviewed by Miller et al.127
The corresponding reaction of methyl radicals with oxygen has also received
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

considerable attention. In the reaction of oxygen with larger alkyl radicals more
complex mechanisms are possible as described recently by Taatjes112 which are not
available. The analogous channels to reaction (2) are:
CH3 + O2 - CH3O + O (R 24a)
- CH3O2 (R 24b)
However, there is also the possibility of another channel forming OH and HCHO
playing a role:
CH3 + O2 - CH2O + OH (R 24c)
159
Using a narrow width, cw, ring dyelaser, Herbon et al. observed the production of
OH from reaction (24c) via UV absorption and O atoms from reaction (24a) via
atomic resonance absorption. They found that the OH channel was dominant over
the experimental temperature range of 16002400 K, in contrast to an earlier
measurement of Michael et al.160 who could find no evidence for channel (24c). It
is thought that the reaction takes place via a curve crossing mechanism to the
CH2O + OH surface (which correlates to electronically excited oxygen).161

5. Multi-mechanism reactions
5.1 Introduction
Sections 3 and 4 considered reactions where the dierent products are formed via
similar mechanisms, e.g. there is a competition between abstraction at two sites or
the dissociation of a complex into two sets of products. In this section we will
consider a number of competitions between dierent mechanisms, but here we note
that competition does not need to be between two dierent types of chemical
reaction. For electronically or vibrationally excited species, the competition can be
between physical deactivation processes and chemical reactions and we start our
discussion with some topical examples in this area.

5.2 Chemical reaction or relaxation


5.2a 1CH2 reactions. Methylene is an important intermediate in both combus-
tion and planetary atmospheres. The ground state of methylene is triplet with the
first singlet state lying only 38 kJ mol!1 higher in energy.162 Despite the similarities
in energies, the reactivities of the two states are significantly dierent, with singlet
methylene generally reacting close to the gas kinetic limit whilst triplet methylene
rate coecients are typically several orders of magnitude less. The coupling of the
close proximity in energies with widely diering reactivity has ensured that there
have been many investigations in energy transfer processes in methylene.
Singlet methylene can conveniently be generated from the photolysis of ketene at
the 308 nm excimer wavelength. In contrast photolysis at 351 nm, another common
excimer wavelength solely generates the electronic ground state. The reactions of
singlet methylene with the rare gases can only lead to relaxation. The relaxation
mechanism is thought to occur via state crossing at certain gateway or mixed
singlet/triplet states followed by rotational relaxation of the triplet state.163,164 Such
processes are characterised by a positive temperature dependence due to increased

206 | Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

population of the mixed states and to the temperature dependence of rotational


relaxation. There is also a clear increase in eciency with the size of the inert
collider.165
Of more interest are reactions with molecular species where now both relaxation
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

and reaction can compete. The overall rate coecient of many 1CH2 reactions
decreases with increasing temperature. If collision induced intersystem crossing
(CIISC) has a positive temperature dependence for these molecular systems, then
the fraction of relaxation will increase substantially with temperature and could
become the dominant process under combustion conditions. The first example
considered is that of 1CH2 with H2 with the reactive pathway leading to the
generation of CH3 + H.
1
CH2 + H2 - CH3 + H (R 25a)
1 3
CH2 + H2 - CH2 + H2 (R 25b)
1
The rate coecient for CH2 by H2 at 298 K does not follow the expected trend in
terms of the relaxation process, suggesting that reaction is important. Reaction can
proceed via the insertion into the H2 bond forming an excited methane intermediate
(Section 4). Redissociation can lead back to methylene in either electronic state, but
the forward channel to product has a lower exit barrier.
The most obvious way to probe the channel eciency of reaction (25) would be to
monitor both singlet and triplet methylene. However, no convenient single system
exists to selectively observe both electronic states. However, it is possible to monitor
the channel eciency by following the production of H by vuvLIF. Channel (25b) is
dark with respect to H atom production, but 3CH2 reacts eciently with NO to
generate H atoms with a known yield.166
The study took place in three stages (Fig. 17) each with identical photolysis
conditions to ensure that the same [1CH2] was being generated in each case. Firstly,
the reaction was carried out in H2 and the H atom signal (J in Fig. 17a) recorded.
Sucient H2 was added such that the reaction was eectively instantaneous on the
timescale of the monitoring; H atoms subsequently decay by diusional loss.
Secondly, a small concentration of NO is added, titrating any 3CH2 to H, resulting
in a slight increase in the H atom signal over the first experiment ( in Fig. 17a).

Fig. 17 (a) Time profiles for experiments 1,2 and 3: O = experiment 1; = experiment 2;
and n = experiment 3. The full line represents the best fit to the data. The traces were recorded
at 20 Torr total pressure and at 295 K. (b) Three experiment analysis at 15 Torr total pressure
and 500 K: O = experiment 1; = experiment 2, and D = experiment 3. Note that there is
little dierence between traces for experiments 1 and 2 at all times, and between traces for
experiments 1, 2 and 3 at long-times, demonstrating the very small channel yield for reaction
25b at this higher temperature. Reproduced with permission from ref. 167.

Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222 | 207


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

Finally the hydrogen was replaced with helium ensuring that all 1CH2 was converted
to 3CH2 and the latter was again converted to H via reaction with NO (m in Fig.
17a). The good agreement in the final signal levels from the second and third
experiments demonstrates that all of the initial methylene has been eectively
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

accounted for.
From Fig. 17a, it is clear to see that the majority of the 1CH2 reacts with the
hydrogen to generate H, and that only a small amount (B15%) is collisionally
relaxed, in good agreement with a previous, indirect, study.168 However, Fig. 17b
demonstrates that the CIISC has a negative temperature dependence for this
reaction; at 500 K, addition of NO to the H2 mixture has a negligible eect on
the amount of H generated, showing that at this temperature virtually all the 1CH2 is
removed via reaction.
Whilst such an observation is incompatible with the gateway model, it is
consistent with a curve crossing mechanism. 1CH2 reacts with H2 without a barrier,
but the 3CH2 reaction has a significant barrier and therefore a curve crossing process
is feasible. As the temperature increases the rate at which the system passes through
the crossing increasing, decreasing the probability for curve crossing, favouring the
formation of the methane intermediate and subsequent dissociation to the CH3 + H
products.
The second example involves the reaction of 1CH2 with oxygen. This reaction has
many thermodynamically accessible channels and 1CH2 reacts with other open shell
species such as NO with a rate coecient close to gas kinetic. However, perhaps
surprisingly, two separate studies have shown that despite the availability of spin and
thermodynamically accessible products, the predominant reaction is physical relaxa-
tion.
Blitz et al. monitored the H atom product from the reaction of 1CH2 with H2 (the
calibration reaction with an H atom yield of 0.850.90168) and with O2 under low
pressure conditions in the absence of a bath gas.169 Fig. 18 shows the rapid, almost
instantaneous production of H atom signal from the calibration reaction. Fig. 18

Fig. 18 Time scales for the reaction of 1CH2 with (a) hydrogen and (b) oxygen. Under the
experimental conditions 1CH2 will be removed with approximately the same time constant in
both experiments, but clearly H atoms are produced on a much longer timescale from the
reaction of 1CH2 with O2. Figure reproduced with permission from ref. 169.

208 | Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

also shows H atom production from reaction of 1CH2 with O2, but the timescale of
the process is very much longer. Rate coecients for 1CH2 removal by H2 and O2 are
comparable (kCH2+H2 = 1.0 ( 10!10 cm3 molecule!1 s!1, kCH2+O2 = 5.2 ( 10!11
cm3 molecule!1 s!1), suggesting that 1CH2 is not the source of the observed H. In
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

fact the risetime of the H atom signal is consistent with production from 3CH2,
suggesting that 100% of the reaction proceeds via collisional deactivation.
Hancock and Haverd (Fig. 19)170 monitored the products of the reaction of
methylene with oxygen via FTIR emission spectroscopy.132 They were able to
observe strong emission from vibrationally excited CO2, H2CO and weaker emission
from CO. The observed product spectrum was independent of the initial state of the
methylene and was produced with the same time dependence, compatible with the
rate coecient for the removal of 3CH2 by O2.171 Additionally, if sucient argon
was added in the presence of O2, to intercept the 1CH2 and relax it to 3CH2 the
intensity of the product emission was essentially unchanged.
The results of Hancock and Haverd contradict some earlier FTIR work where Su
et al.172 observed diering vibrational distributions from the reactions of the singlet
and triplet methylene. However, Hancock and Haverd question the assignment of
emissions near 2000 cm!1 as coming from vibrationally excited CO, and suggest that
these originate either from highly excited CO2 or ketene. Additionally, whilst Su et
al. used a time resolved technique, the concentration of oxygen was such that there is
no temporal resolution of the reaction, which was essentially instantaneous,
although they were able to observe the vibrational relaxation of the products.
The reactions of 1CH2 methylene remain a source of some controversy. The exact
mechanism of CIISC is open to question. Recent studies in this laboratory probing
the rates of removal of ortho and para methylene states suggest that the gateway
model may need refining for reactions with unreactive species. For those reactions
where complex formation is possible either relaxation or chemical reaction can
occur.

Fig. 19 Plot of the pseudo first order rate coecient obtaining from H2CO emission from the
photolysis of CH2CO at 351 nm (100 mTorr, filled squares) and 308 nm (10 mTorr, open
circles) in the presence of O2.170 The straight line is a fit to the 351 nm data, and the slope gives a
rate constant for the 3CH2 + O2 reaction of (3.1 * 0.4) ( 10!12 cm3 molecule!1 s!1, consistent
with previous measurements.171

Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222 | 209


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

5.2b O(1D) reactions. O(1D) is another species where the competition between
removal process can be important. O(1D), produced from ozone photolysis, is a
major source of OH in the atmosphere via reaction with water. In the atmosphere
there is a competition between removal of O(1D) by water and deactivation by
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

collisions with oxygen and nitrogen, but, additionally, there is also a competition
between reaction and relaxation in reaction (26):
O(1D) + H2O - 2OH (R 26a)
1 3
O( D) + H2O - O( P) + H2O (R 26b)
A study by Wine and Ravishankara in 1982, determined that reaction dominated
with the fraction of relaxation being 5 * 3%.173 Recently Carl174 has developed a
very sensitive chemiluminescence method of monitoring both ground and excited
states of oxygen atoms which allowed for a much more precise determination of the
branching ratio for reactions such as reaction (26).
The method made use of the chemiluminescence reaction:
C2H + O(1D) - CH(A2D) + CO (R 27)
3 2
C2H + O( P) - CH(A D) + CO (R 28)
1
C2H and O( D) radicals were generated in the same photolysis pulse and the
chemiluminescence at 431 nm was monitored continuously (giving a significant
improvement in duty cycle over pulsed detection methods). O(1D) reacts much faster
than the triplet state (and C2H) and hence the signal was biexponential in nature,
with the fast decay corresponding to O(1D) removal, the slower decay representing
O(3P) removal. Fig. 20 shows the precision of the technique; signal was detected over
B3 orders of magnitude and clearly the yield O(3P) from reaction (26b) at an upper
limit of 0.3% is much less than the 5% determined in the Wine and Ravishankara
study.

Fig. 20 Two Iem decay profiles taken at 10 Torr total pressure (He). The upper profile is taken
under conditions of [C2H2] = 5.8 ( 1014 cm!3and [N2O] = 3.3 ( 1014 cm!3. The lower profile
under the same conditions as the upper one but with 6.0 ( 1015 cm!3 He replaced by H2O. The
dotted lines are simulations in which various fractions of O(1D2) are quenched by H2O. The fit
to the actual profile (not shown for the sake of clarity) requires no quenching by H2O. Figure
reproduced with permission of PCCP owner societies from ref. 174.

210 | Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

5.3 The reaction of OH with acetone


For the reaction of OH radicals with oxygenated species there is potential for
multiple mechanisms for reaction with addition/elimination reactions potentially
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

competing with abstraction, and for the mechanism of the abstraction reaction to be
complicated by an attractive potential energy surface with possible complex forma-
tion. There has been some debate about the mechanisms of several OH reactions with
oxygenated VOC and the arguments are typified by the reaction of OH with acetone.
In 2000, Wollenhaupt et al.175 reported a kinetic study on the removal of OH by
acetone. The temperature dependence of the rate coecient was non-Arrhenius in
nature as shown in Fig. 21 and Wollenhaupt et al. speculated that one possible
explanation was that the reaction was proceeding via two mechanisms; at high
temperatures abstraction dominated, whereas at lower temperatures a complex is
formed which can dissociate to give acetic acid and methyl radicals.
OH + CH3COCH3 - H2O + CH3COCH2 (R 29a)
OH + CH3COCH3 - complex - CH3 + CH3C(O)OH (R 29b)
Such a hypothesis is of great interest as not only does it suggest that multiple
mechanisms are occurring, it might also account for the significant concentrations of
acetic acid observed in the troposphere.176 Wollenhaupt et al. recognised that their
kinetic data could be interpreted by other explanations and in a follow up paper

Fig. 21 Arrhenius plot of all data obtained using both PLP-RF (J) and PLP-PLIF. The data
are fit (central solid line) by a curve of the form k(t) = A1 exp(!E1/T) + A2 exp(E2/T). The
outermost solid curves represent the 95% confidence limits. The dotted lines are the individual
contributions from the positive and negative temperature-dependent terms in k(t). Represen-
tative error bars are shown for just one PLP-RF (10%) and one PLP-PLIF (5%) data point
close to 300 K for clarity. Figure reproduced with permission from ref. 175.

Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222 | 211


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

obtained more direct information on the products. Neither acetic acid nor methyl
radicals are directly detectable by the pulsed laser induced fluorescence techniques
available to Wollenhaupt and Crowley, who therefore carried out the reactions in
the presence of NO2 to convert methyl radicals into the detectable methoxy
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

species.177

CH3 + NO2 - CH3O + NO (R 30)

The system is relatively complex in that reaction (30) has a pressure dependent
channel forming dark CH3NO2 product and that the methoxy radical will also react
with NO2. Wollenhaupt and Crowley carefully studied the reactions of methyl and
methoxy radicals with NO2 and investigated the sensitivity of their experiments to
the fraction of reaction (29) producing acetic acid, a, using numerical models. They
found that a was 0.50 * 0.15 at 297 K, but interestingly decreased to 0.3 * 0.1 at
233 K in contrast to their expectations from the kinetic studies. Wollenhaupt and
Crowley suggested that this would be consistent with CH3C(O)OH being formed by
decomposition of an alkoxy intermediate; at lower temperatures more of the alkoxy
radical is intercepted by reaction with NO2. Whilst more direct than the kinetic
studies, Wollenhaupt and Crowley recognised that their system was complex and
that there were other routes to methyl radical formation, although their simulations
suggested that such routes were not compatible with the observed timescales of
methoxy radical production.
Supportive evidence for multiple channels in reaction (29) rapidly came from a
complimentary study by Vasvari et al.178 In a discharge flow reactor they monitored
the LIF signal from the acetonyl radical generated in reaction (29a); comparing the
observed signal from the reaction of F + acetone (assumed to proceed 100% via
acetonyl formation) and the signal observed when water was added after the
discharge to convert the F atoms into OH. The fractional yield of acetonyl at room
temperature was 0.50 * 0.04, in good agreement with Wollenhaupt and Crowley.
Vasvari et al. recognised the importance of ab initio calculations in probing the
mechanism of the reaction and carried out relatively low level calculations. These
showed that relatively strongly bound complexes were formed for the addition and
abstraction channels (1525 kJ mol!1, respectively), but that both the abstraction
reaction and the formation of the alkoxy intermediate (which decomposes to give
acetic acid + methyl radical) had positive activation energies. The activation energy
for abstraction was in fact lower than for the addition channel incompatible with
Wollenhaupt et al.s argument of an addition/elimination channel which dominates
at low temperatures. Vasvari et al. therefore suggested that there may be a concerted
mechanism for acetic acid + methyl formation.
These results and their implications generated a host of follow-up studies, the first
of which was a discharge flowtube/mass spectrometric study by Vandenberk et al.179
OH radicals were generated by reacting a known concentration of NO2 with excess
H atoms from the discharge. Under these conditions the concentration of OH is
equal to the known concentration of NO2. The OH was reacted with excess acetone
and the acetic acid signal (m/z = 60, previously calibrated at the appropriate levels
with known standards) was monitored. In contrast to the earlier experiments, no
evidence was found for the significant production of acetic acid and an upper limit of
5% was placed on channel (29a).
Another negative result for acetic acid formation was reported by Tyndall et al.180
Acetic acid has a strong characteristic IR absorption making it an ideal target for
study in chambers equipped with FTIR spectrometers. Fig. 22 shows an IR spectrum
before irradiation of an acetone/OH precursor mix, (A), after irradiation, (B), the

212 | Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222


This journal is '
c The Royal Society of Chemistry 2007
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18. View Article Online

Fig. 22 IR absorbance spectra (base 10) recorded before (A) and after (B) a 43 min irradiation
of a mixture containing 224 mTorr of acetone, 210 mTorr of CH3ONO, 12 mTorr NO, and
8 mTorr C2H4 in 700 Torr of air diluent at 296 K in the Ford chamber. Spectrum (C) shows the
results of subtracting (B) from (A). A reference spectrum of CH3C(O)OH is shown in panel D.
Figure reproduced with permission of PCCP owner societies from ref. 180.

dierence spectrum, (C) and finally a reference spectrum of acetic acid (D). Clearly
no significant amounts of acetic acid have been formed and an upper limit of 10%
was reported based on the signal to noise of the dierence spectrum and also the
incomplete carbon balance of the expected products compared to acetone loss.
Comprehensive studies from the group of Ravishankara at NOAA seem to
confirm that acetic acid is not a significant product of reaction (29).181,182 Talukdar
et al.182 used discharge flow coupled with chemical ionization mass spectrometry
(CIMS) and in contrast to previous studies, were able to monitor the potential
products from both channels using one technique, minimising the potential for
systematic errors. The reaction of Cl atoms with acetone was used as a calibration

Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222 | 213


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

channel for reaction (29a). Cl atoms were generated from the reaction of H atoms
from the discharge with excess Cl2:
H + Cl2 - H + HCl (R 31)
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

Cl + CH3COCH3 - HCl + CH2COCH3 (R 32)


Product yield studies were carried out in back-to-back experiments where Cl2 was
replaced with NO2, with the H atom concentration maintained at a constant level to
ensure that equal concentrations of Cl and OH were generated. Acetonyl radical
production was monitored and acetonyl concentrations from the two reactions were
essentially unity over the temperature range 242351 K. Additionally the CIMS was
tuned to monitor masses characteristic of acetic acid (with mass spectra and
sensitivities determined by calibrations with pure acetic acid); an upper limit of
1% was placed on the branching ratio to acetic acid formation. A similar study
confirming the dominance of the acetonyl channel has also been reported by Turpin
et al.183
Currently the consensus is that reaction (29a) is essentially the sole channel in the
reaction of OH with acetone, although a recent relative rate study by Ra et al.184
does report a low yield (B5%) of acetic acid that cannot be explained by secondary
chemistry. Several groups have reproduced the strong non-Arrhenius behaviour in
reaction (29) and have shown that the strong curvature can be explained without the
need to evoke a change in reaction mechanism.185,186

5.4 Alkoxy radical reactions


A good example of competition in unimolecular reactions occurs in the alkoxy
radicals that are important intermediates in the atmospheric oxidation of hydro-
carbons. Alkoxy radicals are primarily formed from the exothermic reaction of
peroxy radicals with NO.
RO2 + NO - RO + NO2 (R 33)
Indeed this exothermicity has been postulated to generated chemically activated
alkoxy radicals,187 however under atmospheric conditions thermal distributions of
alkoxy radicals predominate. For large alkoxy radicals isomerization can compete
with decomposition and additionally, the unimolecular reactions of alkoxy radicals
are normally in competition with reaction with molecular oxygen if an a-hydrogen is
available. The 2-pentoxy radical is a species where all three reactions can compete
under atmospheric conditions:188
CH3(CH2)2CH(O)CH3 - CH3CHO + n-C3H7; k = 2.2 ( 104 s!1 (R 34a)
4 !1
- (CH2)3CH(OH)CH3; k = 1.7 ( 10 s (R 34b)
4 !1
+O2 - CH3(CH2)COCH3 + HO2; k = 1 ( 10 s (R 34c)
The competition between the various channels is important in terms of the products
formed. Carbonyl species are formed in both channels (34a) and (34c), but the
aldehydes formed in channel (34a) has a much shorter atmospheric lifetime than
the ketone formed in channel (34c). Following isomerization, the new radical
can add oxygen to form a peroxy radical which in turn will react with NO generat-
ing a hydroxy substituted alkoxy radical which lead to more highly oxygenated
carbonyl species than channels (34a) and (34c). Additionally there has been an
extra NO to NO2 conversion with implications in tropospheric ozone formation.
Experimental work on alkoxy radical reactions has been reviewed by Orlando
et al.188

214 | Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

5.5 Addition vs. abstraction


5.5a The reaction of radicals with unsaturated organic species. The reaction of
radicals, X, with hydrocarbons containing multiple bonds can usually take place via
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

either addition or abstraction and the example of the reaction of Cl with propene has
already been mentioned.82
Cl + C3H6 - C3H6Cl (R 7a)
- HCl + C3H5 (R 7b)
In general at relatively low temperatures addition will dominate, but as the
temperature rises abstraction reactions, which tend to have higher A factors and
activation barriers will contribute more and more to the overall loss of the radical.
For larger species, abstraction can place at a variety of sites, but except at very high
temperatures, abstractions forming resonantly stabilized radicals will predominate.
We note that it is possible for the products of the abstraction to be formed by an
alternative addition elimination reaction. Such reactions are likely to occur via the
formation of a long lived complex with statistical distribution of the energy above
the dissociation barrier to products.189 The vibrational distribution of the HX
product is therefore most likely to be statistical in nature, whereas direct abstraction
with a relatively elongated HX bond at the transition state can lead to the
production of a vibrational inversion in HX.190

5.5b The reaction of R + O2. Of fundamental importance in combustion is the


competition between dierent mechanisms in the reactions of alkyl radicals with
oxygen. At relatively low temperatures, a termolecular addition process occurs
forming the RO2. At higher temperatures the lifetime of the RO2 species with respect
to redissociation to products becomes very short giving the well characterized region
of negative temperature dependence in ignition delay experiments.1 However, as the
temperature is increased further then alternative pathways become available either
involving the energized RO2* species or from direct abstraction and formation of
HO2. A very recent review of this topic by Taatjes has already been mentioned.112

5.5c The reaction of Cl atoms with alkyliodides. The competition between


addition and abstraction is demonstrated in this final example. The reaction of Cl
atoms with methyl iodide has been comprehensively studied, monitoring the removal
of Cl atoms in a laser flash photolysis, resonance fluorescence study by Ayhens
\et al.191 At low temperatures (o260 K) addition to form the CH3ICl complex in a
termolecular reaction dominates and pressure dependent bimolecular rate coe-
cients were recorded. As the temperature increases (263309 K), the Cl atom decays
become biexponential in nature as the addition complex (R 35a) starts to re-
dissociate. Ayhens et al. were able to observe the approach to equilibrium and,
from the extracted forward and reverse rate coecients as a function of temperature,
determine DfH and DfS of the adduct. At still higher temperatures (4364 K), the
lifetime of the adduct becomes insignificant and Cl atoms are removed by abstrac-
tion (R 35b)
Cl + CH3I - CH3ICl (R 35a)
- HCl + CH2I (R 35b)
Subsequently the adduct has been observed directly using cavity ring down
spectroscopy by Enami et al. at 250 K.192,193 In agreement with Ayhens et al. they
observe predominantly adduct formation. Very recently Gravestock et al.194 have

Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222 | 215


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

characterized a laser induced fluorescence spectrum for the adduct. In the reaction of
Cl with CH2I2 adduct formation appears to be irreversible; Gravestock et al. suggest
that ICl elimination may be the dominant fate of the adduct in this case.
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

6. Summary
The 70s and 80s could arguably be considered the decades of the overall rate
coecients. A majority of kinetics studies focused on measuring the rate of removal
of radical species; quantitative product studies were limited to end product analysis
where often complex mechanisms were required to link the concentrations of the
measured stable end products to the products of the initial elementary study.
Developments in experimental techniques and the recognition from theory and
modelling that product branching ratios were required, stimulated increased direct,
quantitative studies of the products of elementary reactions during the 1990s up to
the present. More recently experimental work has been accelerated by the avail-
ability of IR and mass spectrometric techniques to complement the established UV
and visible laser absorption and laser induced fluorescence techniques. Laval
expansion techniques, high pressure chambers and developments in shock tube
experimentation have enormously expanded the range of conditions under which
kinetic and branching ratio information can be obtained. A further major impetus
has come from the increasing accessibility of computational methods, for calculating
potential energy surfaces and for calculating the rate coecients in barrierless
reactions and across multiple potential wells. Many of the examples considered in
this review have been illustrated with potential energy surfaces calculated in
conjunction with the experimental work. Obviously this is of great benefit to the
reader although the danger of treating such calculations as a black box and
applying them inappropriately needs to be recognised.
Looking to the future, the accessibility of theory will be further increased as more
programmes are made user friendly and the increasing power of the PC allows
calculations to be routinely performed at the desk top. Developments in computer
power are also allowing the development of more comprehensive models for atmo-
spheric chemistry (for other planetary atmospheres as well as Earth) and combus-
tion. It is now possible to conceive that in the not too distant future, relatively
complex chemical mechanisms, requiring information on individual elementary
reactions and their products, can be coupled to transport models for real world
application.
In the experimental fields the universal nature of mass spectrometric detection
(especially coupled with the multiplex advantage of time-of-flight techniques) looks
set to increase the amount of data on product branching ratios. The advantages of
IR detection in terms of specificity and providing dynamic information have long
been recognised, but relatively low sensitivity and instrumental reliability and
expense have been prevented full exploitation of these advantages. Quantum cascade
lasers or frequency mixing techniques could open up this field of research. Finally,
the role of chamber experiments should not be ignored. Many important reactions
are not amenable to direct, isolated study via flash photolysis or discharge flow. The
diculties of extracting precise information from stable products possibly formed
several steps after the target reactions have already been mentioned. However,
chambers such as EUPHORE195 and SAPHIR196 have been developed to allow
direct detection of radical species providing further important constraints on the
mechanisms and allowing for more precise information on product branching ratios.

216 | Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

Atmospheric and combustion chemistry have been two of the main practical
drivers for kinetics work. These are now very mature subjects but there are still
important challenges to consider. The increasing use of biofuels as neat fuels,
blended with fossil fuel or processed to provide hydrogen or methanol for fuel cells
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

provides new areas for work with increased focus on oxygenated species. In the field
of atmospheric chemistry there is increasing recognition of role of biogenic emissions
and in the formation and growth of aerosols from relatively involatile atmospheric
oxidation products. Environmental concerns centered around stratospheric ozone
depletion, acid rain formation and air quality have stimulated much of the growth in
kinetics during the 1970s1990s. Issues related to climate change will maintain or
enhance the importance of reaction kinetics for the foreseeable future.

References
1 M. J. Pilling, S. H. Robertson and P. W. Seakins, in Low-temperature combustion and
auto-ignition, ed. M. J. Pilling, Amsterdam, 1997.
2 G. Hancock, L. Lanyi, J. P. Sucksmith and B. K. Woodcock, Pure and Applied
Chemistry, 1994, 66, 1207.
3 M. N. R. Ashfold, P. W. May, J. R. Petherbridge, K. N. Rosser, J. A. Smith, Y. A.
Mankelevich and N. V. Suetin, Physical Chemistry Chemical Physics, 2001, 3, 3471.
4 R. Atkinson and J. Arey, Chemical Reviews, 2003, 103, 4605.
5 G. P. Smith and D. Nash, Icarus, 2006, 182, 181.
6 J. I. Moses, B. Bezard, E. Lellouch, G. R. Gladstone, H. Feuchtgruber and M. Allen,
Icarus, 2000, 143, 244.
7 I. W. M. Smith, Angewandte Chemie-International Edition, 2006, 45, 2842.
8 N. Choi, M. A. Blitz, K. W. McKee, M. J. Pilling and P. W. Seakins, Chemical Physics
Letters, 2004, 384, 68.
9 H. Hippler, N. Krasteva, S. Nasterlack and F. Striebel, Journal of Physical Chemistry A,
2006, 110, 6781.
10 R. X. Fernandes, K. Luther and J. Troe, Journal of Physical Chemistry A, 2006, 110,
4442.
11 A. Fontijn, Pure and Applied Chemistry, 1998, 70, 469.
12 I. W. M. Smith, I. R. Sims and B. R. Rowe, Chemistry-a European Journal, 1997, 3, 1925.
13 I. W. M. Smith, Chemical Reviews, 2003, 103, 4549.
14 D. B. Atkinson and M. A. Smith, Review of Scientific Instruments, 1995, 66, 4434.
15 T. Spangenberg, S. Kohler, B. Hansmann, U. Wachsmuth, B. Abel and M. A. Smith,
Journal of Physical Chemistry A, 2004, 108, 7527.
16 A. B. Vakhtin, S. Lee, D. E. Heard, I. W. M. Smith and S. R. Leone, Journal of Physical
Chemistry A, 2001, 105, 7889.
17 Vohringer-Martinez, B. Hansmann, H. Hernandez, J. S. Francisco, J. Troe and B. Abel,
Science (Washington, D. C., 1883-), 2007, 315, 497.
18 I. W. M. Smith, Science (Washington, D. C., 1883-), 2007, 315, 470.
19 S. Lee and S. R. Leone, Chemical Physics Letters, 2000, 329, 443.
20 J. T. Herbon, R. K. Hanson, C. T. Bowman and D. M. Golden, Proceedings of the
Combustion Institute, 2005, 30, 955.
21 M. A. Oehlschlaeger, D. F. Davidson and R. K. Hanson, Proceedings of the Combustion
Institute, 2005, 30, 1119.
22 M. A. Oehlschlaeger, D. F. Davidson and R. K. Hanson, Journal of Physical Chemistry
A, 2006, 110, 9867.
23 S. H. Song, R. K. Hanson, C. T. Bowman and D. M. Golden, Journal of Physical
Chemistry A, 2002, 106, 9233.
24 J. V. Michael, Progress in Energy and Combustion Science, 1992, 18, 327.
25 H. Umemoto, N. Hachiya, E. Matsunaga, A. Suda and M. Kawasaki, Chemical Physics
Letters, 1998, 296, 203.
26 K. Hitsuda, K. Takahashi, Y. Matsumi and T. J. Wallington, Journal of Physical
Chemistry A, 2001, 105, 5131.
27 J. S. Pilgrim, R. T. Jennings and C. A. Taatjes, Review of Scientific Instruments, 1997, 68,
1875.
28 M. H. Harwood, D. M. Rowley, R. A. Cox and R. L. Jones, Journal of Physical
Chemistry A, 1998, 102, 1790.

Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222 | 217


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

29 D. M. Rowley, W. J. Bloss, R. A. Cox and R. L. Jones, Journal of Physical Chemistry A,


2001, 105, 7855.
30 J. D. DeSain, P. Y. Hung, R. I. Thompson, G. P. Glass, G. Scuseria and R. F. Curl,
Journal of Physical Chemistry A, 2000, 104, 3356.
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

31 R. J. Hoobler and S. R. Leone, Journal of Physical Chemistry A, 1999, 103, 1342.


32 G. He, I. Tokue, L. B. Harding and R. G. MacDonald, Journal of Physical Chemistry A,
1998, 102, 7653.
33 N. Choi, M. J. Pilling, P. W. Seakins and L. Wang, Physical Chemistry Chemical Physics,
2006, 8, 2172.
34 J. P. Meyer and J. F. Hershberger, Journal of Physical Chemistry B, 2005, 109, 8363.
35 Y. D. Gao and R. G. MacDonald, Journal of Physical Chemistry A, 2006, 110, 977.
36 A. A. Kosterev and F. K. Tittel, Ieee Journal of Quantum Electronics, 2002, 38, 582.
37 C. A. Taatjes and J. F. Hershberger, Annual Review of Physical Chemistry, 2001, 52, 41.
38 J. Wang, H. L. Chen, G. P. Glass and R. F. Curl, Journal of Physical Chemistry A, 2003,
107, 10834.
39 L. E. Christensen, M. Okumura, S. P. Sander, R. R. Friedl, C. E. Miller and J. J. Sloan,
Journal of Physical Chemistry A, 2004, 108, 80.
40 G. E. Hall and S. W. North, Annual Review of Physical Chemistry, 2000, 51, 243.
41 J. J. Scherer, J. B. Paul, A. Okeefe and R. J. Saykally, Chemical Reviews, 1997, 97, 25.
42 M. Marzurenka, A. J. Orr-Ewing, R. Peverall and G. A. Ritchie, Annual Reports Section
C, 2005, 101, 100.
43 S. S. Brown, A. R. Ravishankara and H. Stark, Journal of Physical Chemistry A, 2000,
104, 7044.
44 I. R. Slagle, D. Gutman, J. W. Davies and M. J. Pilling, Journal of Physical Chemistry,
1988, 92, 2455.
45 C. Fockenberg, H. J. Bernstein, G. E. Hall, J. T. Muckerman, J. M. Preses, T. J. Sears
and R. E. Weston, Review of Scientific Instruments, 1999, 70, 3259.
46 N. Hansen, S. J. Klippenstein, C. A. Taatjes, J. A. Miller, J. Wang, T. A. Cool, B. Yang,
R. Yang, L. X. Wei, C. Q. Huang, F. Qi, M. E. Law and P. R. Westmoreland, Journal of
Physical Chemistry A, 2006, 110, 3670.
47 C. A. Taatjes, S. J. Klippenstein, N. Hansen, J. A. Miller, T. A. Cool, J. Wang, M. E.
Law and P. R. Westmoreland, Physical Chemistry Chemical Physics, 2005, 7, 806.
48 M. A. Blitz, M. T. Baeza Romero and M. J. Pilling, Review of Scientific Instruments,
2007, in press.
49 A. Fernandez-Ramos, J. A. Miller, S. J. Klippenstein and D. G. Truhlar, Chemical
Reviews, 2006, 106, 4518.
50 L. A. Curtiss, P. C. Redfern, K. Raghavachari and J. A. Pople, Journal of Chemical
Physics, 2001, 114, 108.
51 B. J. Lynch, Y. Zhao and D. G. Truhlar, Journal of Physical Chemistry A, 2005, 109,
1643.
52 W. V. Doering and X. Zhao, Journal of the American Chemical Society, 2006, 128, 9080.
53 S. J. Klippenstein, Y. Georgievskii and L. B. Harding, Physical Chemistry Chemical
Physics, 2006, 8, 1133.
54 J. A. Miller and B. C. Garrett, International Journal of Chemical Kinetics, 1997, 29, 275.
55 J. A. Miller and S. J. Klippenstein, International Journal of Chemical Kinetics, 1999, 31,
753.
56 A. F. Wagner and J. M. Bowman, Journal of Physical Chemistry, 1987, 91, 5314.
57 M. J. Pilling and S. H. Robertson, Annual Review of Physical Chemistry, 2003, 54, 245.
58 J. A. Miller and S. J. Klippenstein, Journal of Physical Chemistry A, 2006, 110, 10528.
59 J. R. Barker and D. M. Golden, Chemical Reviews, 2003, 103, 4577.
60 V. D. Knyazev and W. Tsang, Journal of Physical Chemistry A, 2000, 104, 10747.
61 H. B. Singh, Y. Chen, A. Staudt, D. J. Jacob, D. R. Blake, B. Heikes and J. Snow, Nature
(London), 2001, 410, 1078.
62 P. Cadman, A. W. Kirk and A. F. Trotman-Dickenson, Journal of the Chemical Society,
Faraday Transactions I, 1976, 72, 1027.
63 D. Sarzynski and B. Sztuba, International Journal of Chemical Kinetics, 2002, 34, 651.
64 H. B. Qian, D. Turton, P. W. Seakins and M. J. Pilling, International Journal of Chemical
Kinetics, 2002, 34, 86.
65 G. S. Tyndall, J. J. Orlando, T. J. Wallington, M. Dill and E. W. Kaiser, International
Journal of Chemical Kinetics, 1997, 29, 43.
66 P. A. Hooshiyar and H. Niki, International Journal of Chemical Kinetics, 1995, 27, 1197.
67 S. M. Aschmann and R. Atkinson, International Journal of Chemical Kinetics, 1995, 27,
613.

218 | Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

68 E. Tschuikow-Roux, T. Yano and J. Niedzielski, Journal of Chemical Physics, 1985, 82,


65.
69 C. A. T4aatjes, L. K. Christensen, M. D. Hurley and T. J. Wallington, Journal of Physical
Chemistry A, 1999, 103, 9805.
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

70 C. A. Taatjes, Chemical Physics Letters, 1999, 306, 33.


71 S. Rudic, C. Murray, D. Ascenzi, H. Anderson, J. N. Harvey and A. J. Orr-Ewing,
Journal of Chemical Physics, 2002, 117, 5692.
72 P. W. Seakins, J. J. Orlando and G. S. Tyndall, Physical Chemistry Chemical Physicis,
2004, 6, 2224.
73 A. Persky, Journal of Chemical Physics, 1973, 59, 5578.
74 D. C. Robie, S. Arepalli, N. Presser, T. Kitsopoulos and R. J. Gordon, Chemical Physics
Letters, 1987, 134, 579.
75 R. K. Talukdar, T. Gierczak, L. Goldfarb, Y. Rudich, B. S. Madhava Rao and A.R.
Ravishankara, Journal of Physical Chemistry, 1996, 100, 3037.
76 T. Laurent, P. D. Naik, H. R. Volpp, J. Wolfrum, T. Arusi-Parpar, I. Bar and S.
Rosenwaks, Chemical Physics Letters, 1995, 236, 343.
77 N. Presser and R. J. Gordon, Journal of Chemical Physics, 1985, 82, 1291.
78 A. Persky, Chemical Physics Letters, 2005, 401, 455.
79 R. Schinke and W. A. Lester, Journal of Chemical Physics, 1980, 72, 3754.
80 M. K. Gilles, J. B. Burkholder, T. Gierczak and A. R. Ravishankara, Journal of Physical
Chemistry, 2002, 106A, 5358.
81 W. P. Hess and F. P. Tully, Journal of Physical Chemistry, 1989, 93, 1944.
82 J. S. Pilgrim and C. A. Taatjes, Journal of Physical Chemistry A, 1997, 101, 5776.
83 S. Hatakeyama, N. Washida and H. Akimoto, Journal of Physical Chemistry, 1986, 90,
173.
84 L. Y. Yeung, M. J. Pennino, A. M. Miller and M. J. Elrod, Journal of Physical Chemistry
A, 2005, 109, 1879.
85 R. Volkamer, L. T. Molina, M. J. Molina, T. Shirley and W. H. Brune, Geophysical
Research Letters, 2005, 32, L08806.
86 B. Chuong and P. S. Stevens, Journal of Physical Chemistry A, 2000, 104, 5230.
87 M. Francisco-Marquez, J. R. Alvarez-Idaboy, A. Galano and A. Vivier-Bunge, Physical
Chemistry Chemical Physics, 2003, 5, 1392.
88 I. R. Slagle, D. Gutman, J. W. Davies and M. J. Pilling, Journal of Physical Chemistry,
1988, 92, 2455.
89 A. F. Wagner and D. M. Wardlaw, Journal of Physical Chemistry, 1988, 92, 2462.
90 Y. Bedjanian and G. Poulet, Chemical Reviews, 2003, 103, 4639.
91 R. S. Zhu and M. C. Lin, Journal of Chemical Physics, 2003, 118, 4094.
92 D. M. Golden, International Journal of Chemical Kinetics, 2003, 35, 206.
93 W. J. Bloss, S. L. Nickolaisen, R. J. Salawitch, R. R. Friedl and S. P. Sander, Journal of
Physical Chemistry A, 2001, 105, 11226.
94 M. T. Butterfield, T. Yu and M. C. Lin, Chemical Physics, 1993, 169, 129.
95 I. R. Sims, J. L. Queelec, D. Travers, B. R. Rowe, L. B. Herbert, J. Karthauser and I. W.
M. Smith, Chemical Physics Letters, 1993, 211, 461.
96 J. B. Davey, M. E. Greenslade, M. D. Marshall, M. I. Lester and M. D. Wheeler, Journal
of Chemical Physics, 2004, 121, 3009.
97 K. W. McKee, M. A. Blitz, P. A. Cleary, D. Glowacki, M. J. Pilling, P. W. Seakins and L.
Wang, Journal of Physical Chemistry A, 2007, in press.
98 M. A. Blitz, D. G. Johnson, M. Pesa, M. J. Pilling, S. H. Robertson and P. W. Seakins,
Journal of the Chemical Society-Faraday Transactions, 1997, 93, 1473.
99 H. Thiesemann, J. MacNamara and C. A. Taatjes, Journal of Physical Chemistry A, 1997,
101, 1881.
100 P. Fleurat-Lessard, J. C. Rayez, A. Bergeat and J. C. Loison, Chemical Physics, 2002,
279, 87.
101 C. A. Taatjes and S. J. Klippenstein, Journal of Physical Chemistry A, 2001, 105,
8567.
102 K. W. McKee, M. A. Blitz, K. J. Hughes, M. J. Pilling, H. B. Qian, A. Taylor and P. W.
Seakins, Journal of Physical Chemistry A, 2003, 107, 5710.
103 N. Galland, F. Caralp, Y. Hannachi, A. Bergeat and J. C. Loison, Journal of Physical
Chemistry A, 2003, 107, 5419.
104 N. Yamauchi, A. Miyoshi, K. Kosaka, M. Koshi and H. Matsui, Journal of Physical
Chemistry, 1999, 103, 2723.
105 L. B. Harding, Y. Georgievskii and S. J. Klippenstein, Journal of Physical Chemistry A,
2005, 109, 4646.

Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222 | 219


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

106 P. W. Seakins, S. H. Robertson, M. J. Pilling, D. M. Wardlaw, F. L. Nesbitt, R. P.


Thorn, W. A. Payne and L. J. Stief, Journal of Physical Chemistry, 1997, 101, 9974.
107 M. Brouard, M. T. Macpherson and M. J. Pilling, Journal of Physical Chemistry, 1989,
93, 4047.
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

108 S. H. Robertson, A. F. Wagner and D. M. Wardlaw, Journal of Chemical Physics, 1995,


103, 2917.
109 R. A. Brownsword, A. Canosa, B. R. Rowe, I. R. Sims, I. W. M. Smith, D. W. A.
Stewart, A. C. Symonds and D. Travers, Journal of Chemical Physics, 1997, 106, 7662.
110 I. W. M. Smith, Journal of the Chemical Society, Faraday Transactions, 1997, 93, 3741.
111 K. L. Gannon, D. R. Golwacki, M. A. Blitz, K. J. Hughes, M. J. Pilling and P. W.
Seakins, Journal of Physical Chemistry A, 2007, in press.
112 C. A. Taatjes, Journal of Physical Chemistry A, 2006, 110, 4299.
113 J. V. Michael, D. G. Keil and R. B. Klemm, Journal of Chemical Physics, 1985, 83, 1630.
114 G. S. Tyndall, J. J. Orlando, T. J. Wallington and M. D. Hurley, International Journal of
Chemical Kinetics, 1997, 29, 655.
115 M. A. Blitz, D. E. Heard and M. J. Pilling, Chemical Physics Letters, 2002, 365, 374.
116 M. A. Blitz, D. E. Heard and M. J. Pilling, Journal of Physical Chemistry A, 2006, 110,
6742.
117 M. T. Baeza Romero, M. A. Blitz, D. E. Heard, M. J. Pilling, B. Price, P. W. Seakins and
L. Wang, Faraday Discussions, 2005, 130, 79.
118 P. Devolder, S. Dusanter, B. Lemoine and C. Fittschen, Chemical Physics Letters, 2005,
417, 154.
119 D. Townsend, S. A. Lahankar, S. K. Lee, S. D. Chambreau, A. G. Suits, X. Zhang, J.
Rheinecker, L. B. Harding and J. M. Bowman, Science (Washington, D. C., 1883-), 2004,
306, 1158.
120 I. R. Slagle, D. Sarzynski and D. Gutman, Journal of Physical Chemistry, 1987, 91, 4375.
121 R. Zellner, D. Hartmann, J. Karthauser, P. Rhasa and G. Weilbring, Journal of the
Chemical Society, Faraday Transactions 2, 1988, 84, 549.
122 P. W. Seakins and S. R. Leone, Journal of Physical Chemistry, 1992, 96, 4478.
123 C. Fockenberg, G. E. Hall, J. M. Preses, T. J. Sears and J. T. Muckerman, Journal of
Physical Chemistry A, 1999, 103, 5722.
124 J. M. Preses, C. Fockenberg and G. W. Flynn, Journal of Physical Chemistry A, 2000,
104, 6758.
125 T. P. Marcy, R. R. Diaz, D. E. Heard, S. R. Leone, L. B. Harding and S. Klippenstein,
Journal of Physical Chemistry, 2001, 105A, 8361.
126 D. C. Fang, L. B. Harding, S. J. Klippenstein and J. A. Miller, Faraday Discussions, 2001,
119, 207.
127 J. A. Miller, M. J. Pilling and J. Troe, Proceedings of the Combustion Institute, 2005, 30,
43.
128 J. Peeters, W. Boullart and I. Langhans, International Journal of Chemical Kinetics, 1994,
26, 869.
129 L. B. Harding and A. F. Wagner, Journal of Physical Chemistry, 1986, 90, 2974.
130 S. J. Klippenstein, J. A. Miller and L. B. Harding, Proceedings of the Combustion
Institute, 2002, 29, 1209.
131 P. Zou and D. L. Osborn, Physical Chemistry Chemical Physics, 2004, 6, 1697.
132 P. W. Seakins, in Fourier Transform Infrared Emission Spectroscopy as a Tool for the
Study of Chemical Reactions, eds. K. Liu and A. Wagner, Sinapore, 1995.
133 E. Herbst, Annual Review of Physical Chemistry, 1995, 46, 27.
134 N. Haider and D. Husain, Journal of the Chemical Society, Faraday Transactions 2, 1993,
89, 7.
135 D. Chastaing, S. D. Le Picard, I. R. Sims, I. W. M. Smith, W. D. Geppert, C. Naulin and
M. Costes, Chemical Physics Letters, 2000, 331, 170.
136 R. I. Kaiser, A. M. Mebel, A. H. H. Chang, S. H. Lin and Y. T. Lee, Journal of Chemical
Physics, 1999, 110, 10330.
137 J. C. Loison and A. Bergeat, Physical Chemistry Chemical Physicis, 2004, 6, 5396.
138 A. M. Mebel, R. I. Kaiser and Y. T. Lee, Journal of the American Chemical Society, 2000,
122, 1776.
139 S. E. Paulson and J. J. Orlando, Geophysical Research Letters, 1996, 23, 3727.
140 A. R. Rickard, D. Johnson, C. D. McGill and G. Marston, Journal of Physical Chemistry
A, 1999, 103, 7656.
141 S. M. Aschmann, J. Arey and R. Atkinson, Atmospheric Environment, 2002, 36, 4347.
142 J. H. Kroll, J. S. Clarke, N. M. Donahue, J. G. Anderson and K. L. Demerjian, Journal of
Physical Chemistry A, 2001, 105, 1554.

220 | Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

143 I. Faloona, D. Tan, W. Brune, J. Hurst, D. Barket, T. L. Couch, P. Shepson, E. Apel, D.


Riemer, T. Thornberry, M. A. Carroll, S. Sillman, G. J. Keeler, J. Sagady, D. Hooper
and K. Paterson, Journal of Geophysical Research-Atmospheres, 2001, 106, 24315.
144 S. Koch, R. Winterhalter, E. Uherek, A. Kollo, P. Neeb and G. K. Moortgat,
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

Atmospheric Environment, 2000, 34, 4031.


145 G. Marston and D. Johnson, Chemical Society Reviews, 2007, in preparation.
146 S. M. Aschmann, J. Arey and R. Atkinson, Atmospheric Environment, 1996, 30, 2939.
147 S. E. Paulson, J. D. Fenske, A. D. Sen and T. W. Callahan, Journal of Physical Chemistry
A, 1999, 103, 2050.
148 R. Atkinson and S. M. Aschmann, Environmental Science & Technology, 1993, 27, 1357.
149 S. E. Paulson, M. Y. Chung and A. S. Hanson, Journal of Physical Chemistry A, 1999,
103, 8125.
150 P. Neeb and G. K. Moortgat, Journal of Physical Chemistry A, 1999, 103, 9003.
151 R. Gutbrod, S. Meyer, M. M. Rahman and R. N. Schindler, International Journal of
Chemical Kinetics, 1997, 29, 717.
152 J. H. Kroll, N. M. Donahue, V. J. Cee, K. L. Demerjian and J. G. Anderson, Journal of
the American Chemical Society, 2002, 124, 8518.
153 J. H. Kroll, T. F. Hanisco, N. M. Donahue, K. L. Demerjian and J. G. Anderson,
Geophysical Research Letters, 2001, 28, 3863.
154 U. Maas and J. Warnatz, Combustion and Flame, 1988, 74, 53.
155 J. V. Michael, M. C. Su, J. W. Sutherland, J. J. Carroll and A. F. Wagner, Journal of
Physical Chemistry A, 2002, 106, 5297.
156 R. W. Bates, D. M. Golden, R. K. Hanson and C. T. Bowman, Physical Chemistry
Chemical Physics, 2001, 3, 2337.
157 M. Votsmeier, S. Song, R. K. Hanson and C. T. Bowman, Journal of Physical Chemistry
A, 1999, 103, 1566.
158 J. Hahn, L. Krasnoperov, K. Luther and J. Troe, Physical Chemistry Chemical Physicis,
2004, 6, 1997.
159 J. T. Herbon, R. K. Hanson, C. T. Bowman and D. M. Golden, Proceedings of the
Combustion Institute, 2005, 30, 955.
160 J. V. Michael, S. S. Kummaran and M. C. Su, Journal of Physical Chemistry A, 1999, 103,
5942.
161 S. P. Walch, Chemical Physics Letters, 1993, 215, 81.
162 P. Jensen and P. R. Bunker, Journal of Chemical Physics, 1988, 89, 1327.
163 U. Bley and F. Temps, Journal of Chemical Physics, 1993, 98, 1058.
164 G. Hancock and M. R. Heal, Journal of Physical Chemistry, 1992, 96, 10316.
165 M. N. R. Ashfold, M. A. Fullstone, G. Hancock and G. W. Ketley, Chemical Physics,
1981, 55, 245.
166 M. Fikri, S. Meyer, J. Roggenbuck and F. Temps, Faraday Discussions, 2001, 119, 223.
167 M. A. Blitz, N. Choi, T. Kovacs, M. J. Pilling and P. W. Seakins, International
Symposium on Combustion., 2005, 30, 927.
168 M. A. Blitz, M. J. Pilling and P. W. Seakins, Physical Chemistry Chemical Physicis, 2001,
3, 2214.
169 M. A. Blitz, K. W. McKee, M. J. Pilling and P. W. Seakins, Chemical Physics Letters,
2003, 372, 295.
170 G. Hancock and V. Haverd, Chemical Physics Letters, 2003, 372, 288.
171 D. C. Darwin, A. T. Young, H. S. Johnston and C. B. Moore, Journal of Physical
Chemistry, 1989, 93, 1074.
172 H. Su, W. Mao and F. Kong, Chemical Physics Letters, 2000, 322, 21.
173 P. H. Wine and A. R. Ravishankara, Chemical Physics, 1982, 69, 365.
174 S. A. Carl, Physical Chemistry Chemical Physics, 2005, 7, 4051.
175 M. Wollenhaupt, S. A. Carl, A. Horowitz and J. N. Crowley, Journal of Physical
Chemistry A, 2000, 104, 2695.
176 P. Warneck, Journal of Atmospheric Chemistry, 2005, 51, 119.
177 M. Wollenhaupt and J. N. Crowley, Journal of Physical Chemistry A, 2000, 104, 6429.
178 G. Vasvari, I. Szilagyi, A. Bencsura, S. Dobe, T. Berces, E. Henon, S. Canneaux and F.
Bohr, Physical Chemistry Chemical Physics, 2001, 3, 551.
179 S. Vandenberk, L. Vereecken and J. Peeters, Physical Chemistry Chemical Physics, 2002,
4, 461.
180 G. R. S. Tyndall, J. J. Orlando, T. J. Wallington, M. D. Hurley, M. Goto and M.
Kawasaki, Physical Chemistry Chemical Physics, 2002, 4, 2189.
181 T. Gierczak, M. K. Gilles, S. Bauerle and A. R. Ravishankara, Journal of Physical
Chemistry A, 2003, 107, 5014.

Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222 | 221


This journal is '
c The Royal Society of Chemistry 2007
View Article Online

182 R. K. Talukdar, T. Gierczak, D. C. McCabe and A. R. Ravishankara, Journal of Physical


Chemistry A, 2003, 107, 5021.
183 E. Turpin, C. Fittschen, A. Tomas and P. Devolder, Journal of Atmospheric Chemistry,
2003, 46, 1.
Published on 11 April 2007. Downloaded by Universidad Autonoma de Nuevo Leon on 26/01/2017 22:28:18.

184 J. D. Ra, P. S. Stevens and R. A. Hites, Journal of Physical Chemistry A, 2005, 109,
4728.
185 F. Caralp, W. Forst, E. Henon, A. Bergeat and F. Bohr, Physical Chemistry Chemical
Physics, 2006, 8, 1072.
186 T. Yamada, P. H. Taylor, A. Goumri and P. Marshall, Journal of Chemical Physics, 2003,
119, 10600.
187 J. J. Orlando, G. S. Tyndall, L. Vereecken and J. Peeters, Journal of Physical Chemistry
A, 2000, 104, 11578.
188 J. J. Orlando, G. S. Tyndall and T. J. Wallington, Chemical Reviews, 2003, 103, 4657.
189 E. Arunan, S. J. Wategaonkar and D. W. Setser, Journal of Physical Chemistry, 1991, 95,
1539.
190 V. R. Morris, F. Mohammad, L. Valdry and W. M. Jackson, Chemical Physics Letters,
1994, 220, 448.
191 Y. V. Ayhens, J. M. Nicovich, M. L. McKee and P. H. Wine, Journal of Physical
Chemistry A, 1997, 101, 9382.
192 S. Enami, T. Yamanaka, S. Hashimoto, M. Kawasaki and K. Tonokura, Journal of
Physical Chemistry A, 2005, 109, 6066.
193 S. Enami, S. Hashimoto, M. Kawasaki, Y. Nakano, T. Ishiwata, K. Tonokura and T. J.
Wallington, Journal of Physical Chemistry A, 2005, 109, 1587.
194 T. Gravestock, M. A. Blitz and D. E. Heard, Physical Chemistry Chemical Physics, 2007,
in press.
195 W. J. Bloss, J. D. Lee, C. Bloss, D. E. Heard, M. J. Pilling, K. Wirtz, M. Martin-Reviejo
and M. Siese, Atmospheric Chemistry and Physics, 2004, 4, 571.
196 J. Bossmeyer, T. Brauers, C. Richter, F. Rohrer, R. Wegener and A. Wahner, Geophy-
sical Research Letters, 2006, 33.

222 | Annu. Rep. Prog. Chem., Sect. C, 2007, 103, 173222


This journal is '
c The Royal Society of Chemistry 2007

Você também pode gostar