You are on page 1of 22

Molecular Elimination of Methyl Formate in Photolysis at 234 nm :

Roaming vs. Transition State-type Mechanism

Meng-Hsuan Chao, Po-Yu Tsai, and King-Chuen Lin*


Department of Chemistry, National Taiwan University, Taipei 106, and Institute of
Atomic and Molecular Sciences, Academia Sinica, Taipei 106, TAIWAN

Pages: 22
Figures: 3
Table: 1

* Author to whom correspondence should be addressed.


Fax: 886-2-23621483
E-mail: kclin@ccms.ntu.edu.tw

1
Abstract

Ion imaging coupled with (2+1) resonance-enhanced multiphoton ionization


(REMPI) technique is employed to probe CO(v”=0) fragments at different
rotational levels following photodissociation of methyl formate (HCOOCH3) at 234
nm. When the rotational level, J”CO, is larger than 24, only a broad translational
energy distribution extending from 0 to 70 kcal/mol with an average energy of about
23 kcal/mol appears. The dissociation process is initiated on the energetic ground
state HCOOCH3 that surpasses a tight transition state along the reaction coordinate
prior to breaking into CO + CH3OH. This molecular dissociation pathway accounts
for the CO fragment with larger rotational energy and large translational energy.
As J”CO decreases, a bimodal distribution of translational energy arises with one
broad component and the other sharp component carrying the average energy of
only 1-2 kcal/mol. The branching ratio of the sharp component increases with
decrease of J”CO; (7.3±0.6)% is reached as the image is probed at J”CO=10. The
production of sharp component is ascribed to a roaming mechanism that has the
following features: a small total translational energy, a low rotational energy
partitioning in CO, but a large internal energy in the CH3OH co-product. The
internal energy deposition in the fragments shows distinct difference from those via
the conventional transition state.

2
I. Introduction

Methyl formate (HCOOCH3) is a byproduct of oxidation of several fuel


alternatives.1-3 It has long been used as a syngas precursor and also applied in the
manufacture of CO with high purity in terms of thermal decomposition. The
theoretical and experimental studies on thermal decomposition of this simplest ester
are actively focused on surfaces with various catalysts4-7 or gas-phase reactions.8-10
The major products observed in gas-phase pyrolysis or shock tube experiments are
methanol (CH3OH), formaldehyde (CH2O), and carbon monoxide (CO).8-10 Steacie8
proposed the following decomposition mechanism. HCOOCH3 is first decomposed
to CH3OH + CO, followed by successive reactions of CH3OHCH2O+H2 and
CH2OCO +H2. Jain and Murwaha9 claimed that the thermal decomposition of
methyl formate is a molecular process without involvement of free radical reaction.
From the theoretical aspects, Francisco11 calculated the probable reaction
mechanism leading to CH3OH+CO, CH4+CO2, CH2O+CH2O, and HCOH+CH2O.
The latter two primary channels provide an alternative pathway for the CH2O
production that is energetically more favorable.
As compared to the thermal decomposition reaction, photodissociation of
methyl formate is much less studied thus far. Their dissociation products essentially
make distinct difference from those by pyrolysis. Pereira and Isolani12 observed the
products of CO, CO2, CH3OH, CH4, and C2H6 following multiphoton infrared
absorption, and suggested two dissociation pathways yielding CH3OH + CO and
CH4 + CO2. Using photofragment translational spectroscopy (PTS), Lee13 reported
four primary dissociation pathways for methyl formate-d (DCOOCH3) at 193.3 nm,
~ ~ ~
leading to the production of CH3O( X 2E)+DCO( X 2A’), CH3O( X 2E)+DCO( A 2A”),
~ ~
CH3OCO( X 2A’)+D(2S), and CH3OD( X 1A’) + CO( X 1  ). The corresponding
branching ratios were determined to be 0.73, 0.06, 0.13 and 0.08, respectively. The
atomic and radical fragments detected above were not found in the thermal
decomposition. The ratio of the molecular dissociation channel to the radical
channel is about 0.08:0.79, whereas the molecular formation of CH3OH + CO is
regarded as the major channel in the pyrolytic process. A typical PTS technique is

3
advantageous for providing data on the kinetic energy distribution and angular
anisotropy in the fragmentation, but lacks the capability of state-resolved detection.
This work is aimed to supplement the information on photodissociation of methyl
formate by detecting the CO fragment at various rotational levels using velocity ion
imaging.
The concept of transition state (TS) has long been considered as foundation of
chemical reaction and unimolecular dissociation. For reactions with potential
barrier, TS is the transient structure at the highest point of the minimum energy
path (MEP) connecting reactants to products. For barrierless reactions, the TS
dividing surface along the reaction coordinate is so selected as to minimize the
reactive flux. The reaction rate coefficients, available energy disposals, and reaction
mechanisms are closely related to this structure. However, recent studies have found
that reactants may bypass TS forming products through non-MEP reactions.14-17
For instance, the dominant reactive channel of F- + CH3COOH→HF + CH2O + OH-
via an ECO2 mechanism was found to be away from intrinsic reaction coordinate.14
In the reaction scheme, H + HBr → H2 + Br, the products with internal energy
exceeding the kinematic limit were expected to arise from reactive trajectories that
deviated significantly from MEP.15
Among these non-MEP reactions, “roaming mechanism” is a new pathway for
unimolecular decomposition.18-20 The mechanism is so termed that recoiling atom or
radical fragments which are weakly bound to the other moiety can meander and
undergo intramolecular abstraction forming molecular products. It usually happens
when the dissociation threshold of a radical channel along a barrierless potential
energy surface (PES) is close to the conventional TS for a molecular channel.
Photodissociation of formaldehyde is the first example ever reported.19-25 Thus far,
only a few molecules are found to decompose through this mechanism including
acetaldehyde,26-31 acetone,32-34 nitrate radical,35 and allyl radical.36 This work
provides the first example in the ester family that involves roaming dynamics
following ultraviolet photodissociation.
Methyl formate is dissociated at 234 nm in this work, followed by detection of
CO fragment using velocity ion imaging coupled with (2+1) resonance-enhanced

4
multiphoton ionization (REMPI) technique. Different rotational levels of CO are
selected for imaging acquisition from which the translational and internal energy
disposals in the fragments are evaluated. Accordingly, two reaction pathways are
found with distinctly different energy partition. One is predicted by the
conventional TS theory and the other is proposed to follow the roaming mechanism.

II. Experimental

The photofragment velocity imaging apparatus similar to that by Eppink


and Parker has been described elsewhere.37-41 It consisted of a source chamber and a
main chamber, both of which were pumped to a low background pressure of ~2×10 -7
Torr. Methyl formate (99.9%) was used as purchased without further purification.
The sample was premixed at ~5% with helium, carried at a total pressure of 40 psi
through a pulsed valve (General Valve Co.) with 270 μs duration and 0.6-mm
diameter orifice operating synchronously with the laser pulses at 10 Hz, and
expanded into the source chamber. After passing through a 1-mm diameter skimmer
and a collimator, the molecular beam was intersected perpendicularly by a linearly
polarized laser beam in a two-stage ion lens region. The skimmer was mounted 30
mm downstream from the nozzle to divide the source chamber from the main
chamber. The electrostatic ion lenses installed in the main chamber consisted of a set
of repeller, extractor, and ground electrodes, each with a central hole of 2, 16, and 16
mm-diameter, respectively.39-41
A Nd:YAG laser (Quanta-Ray, Spectra-Physics) pumped dye laser (ScanMate,
Lambda Physik) operating at 10 Hz was used as photolysis source. Its output was
frequency-doubled to emit at 234 nm with the energy of 80-270 μJ/pulse (BBO-B
crystal, InRad). The radiation was linearly polarized perpendicular to the flight
tube and parallel to the detection surface, and then softly focused to the leading edge
of the skimmed beam with a 300-mm focal length lens to minimize the cluster
formation. A 308 nm XeCl excimer laser (LPX 200, Lambda Physik) pumped dye
laser beam (PD 3000, Lambda Physik) was guided in the opposite direction to probe
the CO(v”=0) fragments. Its output was frequency-doubled to emit at 230 nm with
the energy of 15-30 μJ/pulse. Following the ester photolysis, the focused probe beam

5
with a 200-mm lens was applied in 30 ns delay to ionize CO(v”=0) by the (2+1)
REMPI technique via the B1Σ+←X1Σ+ two-photon transition.42 Laser pulse powers
were adjusted to optimize the two-laser ion signal while nearly eliminating the
single-laser ion signal.
The resulting CO ions were extracted and accelerated into a 36-cm long
field-free drift tube along the molecular beam direction. The ion-cloud expansion
was mapped onto a two-stage microchannel plate (MCP) and a phosphor screen
(FM3040, Galileo). The MCP could be gated within a minimum duration of 250 ns
for mass selection. The ion imaging on the phosphor screen was recorded by a
charge coupled device (CCD) camera (200XL4078, Pixelfly). All the ion signals
without gate restriction may be acquired by a photomultiplier tube and then
transferred to a transient digitizer for display of the time-of-flight (TOF) mass
spectrum.

III. Results and Discussion


A. Broad component of translational energy distribution

As shown in Fig.1a, the raw ion images of CO(v”=0) fragment are acquired
at different rotational levels, following photodissociation of HCOOCH3 at 234 nm.
Each image was accumulated up to 10000-20000 laser shots. The background
images obtained with either photolyzing or ionizing pulse alone were both
subtracted from the raw image. The obtained image is a two-dimensional projection
of the three-dimensional speed and angular distributions with cylindrical symmetry
around the polarization axis of the photolysis laser. The corresponding
three-dimensional spatial distributions of the fragments are reconstructed by the
basis-set expansion method (BASEX).43
The speed distribution P(v) of the fragment, extracted from the reconstructed
ion image, may be converted to the center-of-mass translational energy distribution
P(ET) as follows,
dv
P ( ET )  P (v) , (1)
dET

6
1 mCO
ET  (mCO  mCH3OH )   vCO
2
, (2)
2 mHCOOCH3

where ET is the total translational energy, mCO , mCH3OH and mHCOOCH3 are the mass

of CO, CH3OH, and HCOOCH3, and vCO is the velocity of the CO fragment. Each
corresponding center-of-mass translational energy distribution contains one broad
P(ET)b and one sharp P(ET)s components especially for smaller J”CO images. They
may be fitted using the following formula:

P(ET) = P(ET)s + P(ET)b


= A‧(ET)a1‧[exp(a2‧ET)] + B‧(ET)b1‧[exp(b2‧ET)] (3)

where A and B are the weighing factors for these two components. Each is expressed
by a Boltzmann-like functions with parameters of ai and bi (i=1 and 2), which are
adjusted within 50% variation in the deconvolution process.
Fig.1b shows the resolved components of the P(ET) distribution of CO. The
diffused image acquired at the rotational line of J”CO=24 yields only one broad
translational energy distribution extending beyond 70 kcal/mol and peaking at
about 11-13 kcal/mol with an average translational energy <ET> of 23.4±0.3
kcal/mol. As J”CO becomes smaller, the broad component exhibits similar feature as
the J”CO=24 result, whereas the other much narrower translational component
appears peaking at about 1-2 kcal/mol.
The absorption spectrum of methyl formate peaks at about 215 nm in the
region of 120-260 nm, in which the absorption portion of 200-260 nm is ascribed to
the S1(n, π*)CO excitation.44-46 As mentioned in Introduction, Lee has observed four
dissociation pathways for methyl formate-d (DCOOCH3) at 193.3 nm using PTS
~
technique.13 Among them, the radical channel producing CH3O( X 2E) and
~
DCO( X 2A’) is expected to initiate on the excited state S1 with an anisotropy
parameter of -0.37 and <ET> of 19.5 kcal/mol, while the molecular fragments of
~
CH3OD( X 1A’) and CO( X 1  ) are formed with <ET> of 29.8 kcal/mol from the
ground state So via internal conversion. Apart from the above pathway, the CO

7
products are contributed additionally by the other two minor sources. One is from
dissociative ionization of DCO, i.e., DCO+ D+ + CO and the other is from
spontaneous decomposition of DCO( A 2A”), the upper Renner-Teller state with
linear geometry and higher electronic energy of 26.6 kcal/mol.47 DCO( A 2A”) is
~
coupled to the highly vibrationally excited state X 2A’ from which D + CO are
dissociated. Such a secondary decomposition is characterized with a broad P(ET)
distribution extending from 0 to 12 kcal/mol and <ET> of 4.9 kcal/mol.13
As compared to the PTS results, the broad P(ET) distributions in Fig.1 with
<ET> from 22 to 24 kcal/mol at different J”CO probed is anticipated to be caused by
~
the molecular dissociation channel of CH3OH( X 1A’) and CO( X 1  ). Methyl
formate decomposes on the So state by surpassing an energy barrier of 68.3 kcal/mol
calculated herein, which is close to 74.8 kcal/mol reported previously.11 The obtained
<ET> is slightly smaller than 29.8 kcal/mol in the 193.3 nm photolysis. This result is
reasonable, because a longer photolysis wavelength is applied in this work. The
feature of a widely broad P(ET) distribution is also consistent with that reported by
Lee.13
For inspecting the probability of the HCO( A 2A”) decomposition, the available
energy Eavl in the photodissociation must be known and it is estimated by,

Eavl  Ehv  E0  Eint , (4)

where Ehv is the photon energy of 122.2 kcal/mol corresponding to the photolysis

wavelength 234 nm; Eo denotes the dissociation energy required to break the
CH3O-CHO bond, which is calculated to be 97.1 kcal/mol (in the next section). Eint

is the internal energy of the parent molecule, considered as zero since the rotational
and vibrational excitations are negligible in a supersonic molecular beam. Eavl is

thus estimated to be 25.1 kcal/mol which is 1.5 kcal/mol in short for population at
the A 2A” state. The CO production from the HCO( A 2A”) decomposition is
negligible in our case. Even if a tiny portion of CO is assumed to result from this
secondary decomposition, the corresponding P(ET) distribution is expected to be
broad.13 The results are against the translational energy distribution given in Fig.1

8
which contains only one broad component.
~
The energy barrier to the dissociation scheme of HCO( X 2A’)  CO + H is
~
17.9 kcal/mol.48 Is it probable for energetic HCO( X 2A’) to undergo spontaneous
decomposition? The available energy of 25.1 kcal/mol gained in the radical channel,
HCOOCH3 → CH3O + HCO, will be partitioned into the translational and internal
states of the fragments. Thus, the internal energy deposited in HCO must be
insufficient to break the H-C bond, not to mention intramolecular vibrational
redistribution (IVR) effect. An alternative pathway for HCO to undergo secondary
dissociation is by absorbing one additional photolyzing photon. Nevertheless, as
shown in Fig.2, the photolysis laser energy dependence of the CO intensity at
J”CO=10 gives rise to a slope of 1.20.1, indicating that the CO fragment is mainly
produced by one-photon absorption of methyl formate at 234 nm. The CO
contribution from the dissociative ionization of HCO is neglected.

B. Sharp component of translational energy distribution

As shown in Fig.1b, the sharp component contained in the bimodal


distribution of CO increases as the image is probed at decreased rotational level.
This channel results in a small translational energy as compared to the broad
component. The energy partitioning between the sharp and broad components are
characterized in Table I, which contains the <ET> results and the internal energy
deposition in the fragments. The ratio to the sharp component also listed in Table I
is defined as the sharp distribution divided by the total distribution. This ratio
increases from nil to (7.30.6)% as J”CO decreases to 10. The sharp distribution was
not observed previously by using PTS technique.13 There are several reasons to be
speculated. First, the data analysis in PTS is not sensitive enough to resolve a slight
difference of the translational energy distributions with the same masses. For
instance, in the gas-phase photodissociation of propionyl chloride (C2H5COCl) at
248 nm,49,50 the PTS method can not resolve the fragment energy distributions of
Cl(2P3/2) and Cl*(2P1/2); the latter one lies in a state higher by 2.52 kcal/mol. Second,
the PTS experiments are not sensitive to detect very slow products like the sharp

9
component restricted to <3 kcal/mol of translational energy, because they do not
scatter far from the beam.32 The detection limit of lower energy in PTS depends on
the beam velocity and the angle used in the analysis. Therefore, the slow component
may have been detected in Lee’s experiments,13 but can not be resolved for clear
identification. In contrast, the ion imaging coupled with REMPI technique has the
advantage to offer the detailed information on translational and angular
distributions of state-resolved fragments.
The sharp distribution can not be attributed to any dissociation pathways as
observed previously. Then, what mechanism does it come from? The sharp
component appears at low J”CO with a small kinetic energy. Such dynamical analogs
were reported previously. For instance, in the photodissociation of formaldehyde at
the dissociation energy slightly above the radical channel of H+HCO, the resultant
broad translational energy distribution at J”CO=40 is correlated to the H2
co-product with vibrational levels from 0 to 4.19,20 Then, as J”CO is decreased, a
bimodal internal energy distribution exists in H2. The additional slow CO product is
formed together with the H2 moiety in vibrational levels from 5 to 7. When J”CO
keeps smaller, the slow CO product rises similar to that at the intermediate J”CO,
while the fast CO component almost diminishes. The accompanying quasiclassical
trajectory calculations21-25 suggested that the fast CO products with low vibrational
levels of H2 proceeds via the conventional transition state, whereas the slow CO
component in conjunction with highly vibrational excitation of H2 follows a so-called
roaming mechanism. The latter pathway involves one H atom nearly detaching from
the radical channel (H + HCO) wanders in the broad attractive space of the
H+HCO surface until it attracts the other H atom yielding vibrationally hot H2 and
rotationally cold CO. Similar roaming mechanisms are found in acetaldehyde26-31
and acetone.32-34 Because of larger sizes for these compounds, the vibrational levels
of the co-products can not be resolved clearly like H2 in formaldehyde
decomposition.
Fig.3 shows the decomposition pathways of the ground state HCOOCH3, in
which several ones are referred to Francisco’s calculations11 by using
CCSD(T)/6-311G(2df,2p)//MP2/6-311G(2df,2p) level of theory. His calculations were

10
aimed to find out molecular dissociation mechanisms to interpret the thermal
decomposition products, mainly CH3OH + CO, and the pathways leading to CH2O
observed experimentally. But his results lack the radical dissociation channel to
CH3O + HCO, the products observed by Lee. For this reason, we conducted similar
calculations by using B3LYP/cc-pVTZ//CCSD(T)/cc-pVTZ level of theory adopted
from Gaussion03. As shown in Fig.3, the obtained heat of reactions for different
dissociation channels are consistent with those by Francisco.11 Among them, the
radical dissociation channel, HCOOCH3  CH3O + HCO, does not suffer from any
exit barrier, but requires 97.1 kcal/mol to break the C-O bond. The calculated result
~ ~
is consistent with that for the scheme DCOOCH3  CH3O( X 2E) + DCO( X 2A’), in
which 98.7 kcal/mol was estimated from the enthalpies of formation for methyl
formate and the fragments.13,51
Like the cases of formaldehyde and acetaldehyde, we anticipate that the slow
translational energy distribution should result from a roaming mechanism, in which
CH3O meanders away from the reaction coordinate and finally undergoes
intramolecular abstraction of the hydrogen atom from HCO forming CH3OH and
CO. The resultant CO fragment lies in a state of low translation and low rotation,
while the CH3OH co-product is highly ro-vibrationally excited. The energy
distributions are distinctly different from those along the MEP reaction coordinate.
For instance, as J”CO is probed at 10, <ET> and internal energy of CH3OH share a
ratio of 1.1% and 95.5%, in contrast to 19.9% and 76.8% for the fast
component.(Table I)
~ ~
The radical dissociation products of CH3O( X 2E) and HCO( X 2A’) are
correlated adiabatically with two singlet (S1, So) and two triplet (T1, T2) states of
HCOOCH3. Lee suggested that the dissociation is initiated on the S1 surface such
that the reaction may proceed rapidly as inferred by a nonzero anisotropy
parameter.13 If all the radical products are produced along the S1 surface, then the
roaming mechanism can not be found. Therefore, we expect there is a small portion
of methyl formate that undergoes radical dissociation after coupling to the So state
via internal conversion. In this manner, the translational energy release is usually
small because of no exit barrier. This condition is required for the roaming

11
dynamics. The radical CH3O may then have a chance to explore around a wide flat
region of the CH3O+ HCO PES. CH3O weakly attracted by the interaction potential
renders the O atom with an unpaired electron to abstract the exposed H atom on
CHO at some long distance. The CO fragment generated by this mechanism will
exhibit the following dynamical features: a small total translational energy release
and small rotational excitation in conjunction with large internal excitation of the
co-fragment.
Another alternative roaming mechanism may be proposed by initiation from
~
the atomic dissociation channel of HCOOCH3CH3OCO( X 2A’)+H(2S). In the
photodissociation of methyl formate-d, Lee has observed the D atom together with
CH3 and CO2, which are originated from spontaneous decomposition of CH3OCO.13
Given the feature of translational energy distribution, the C-D bond was suggested
to break on the T1 surface with an exit barrier after intersystem crossing S1T1.
Despite difference of the photolysis wavelength, the CH3OCO fragment may
probably decompose to CH3 and CO2 like the case by Lee. If so, the roaming
mechanism via this pathway will be blocked. Another possibility is that a small
population of HCOOCH3 in the S1 state is allowed to undergo internal conversion to
the ground state So. Then, it requires an enthalpy of reaction of 99.6 kcal/mol to
~
proceed the atomic dissociation, HCOOCH3CH3OCO( X 2A’)+H(2S).52 Even if this
channel does not suffer from any exit barrier, it is difficult for the H atom to
abstract the CH3O radical from CO, because CH3OCO appears to have a partial
double bond between CH3O and CO. Thus, the roaming mechanism is less probable
to result from the atomic dissociation pathway.
While inspecting the roaming conditions, methyl formate shows an energy
window of 28.8 kcal/mol between the radical channel threshold and the TS of
molecular channel (Fig.3). The energy window is opened wider than those found in
formaldehyde and acetaldehyde. The fraction of molecular products via the
roaming mechanism may be enhanced if such a window becomes narrow, but this
condition is not necessary for roaming occurrence. How to apply photolysis energy
is another problem to be concerned. The photolysis energy adopted in this work is
well above the radical dissociation threshold by 25.1 kcal/mol, but a small portion of

12
the radical products, that may not separate away because of lack of repulsive
potential interaction, follow a roaming dynamics. An analog in acetaldehyde with
large excess energy has been reported.18,27 As acetaldehyde is dissociated at 308 nm
with 9.2 kcal/mol above the radical dissociation threshold, the fraction of reactive
flux via the roaming mechanism shares 15%.18 When the radiation at 248 nm is used
with an excess energy of 31.6 kcal/mol above, the fraction is enhanced by a factor of
1.7 at J”CO=7 of v”=0.27 The photolysis wavelength dependence of roaming channel
is also given in formaldehyde.20 The branching ratio of radical channel increases
rapidly with increasing photolysis energy, whereas the molecular channel drops
with energy. As a result, the branching ratio of roaming channel relative to the
conventional molecular channel grows with energy, but the net roaming yield
remains near flat or slightly decreases as the radical channel is taken into account.

VI. Conclusion

The bimodal translational energy distributions are found in the


photodissociation of methyl formate at 234 nm using ion imaging coupled with (2+1)
REMPI technique. The CO(v”=0) fragment is probed at different rotational
quantum states. As J”CO is large (>24), only a broad component with a large average
translational energy of about 23 kcal/mol appears. This dissociation pathway has
similar features as methyl formate-d in the photolysis at 193.3 nm by using PTS
method. HCOOCH3 is expected to surpass a tight transition state prior to breaking
into CO + CH3OH. The other two minor contributions to CO, dissociative ionization
of HCO and decomposition of excited state HCO( A 2A”), are negligible in this work.
As J”CO decreases, a bimodal distribution of translational energy arises with one
broad component and the other sharp component. The branching ratio of sharp
component increases with the decrease of J”CO. The formation of sharp component
is ascribed to a roaming mechanism which is characterized by a low rotational
energy deposition in CO in conjunction with a small total translational energy, but a
large internal energy in the CH3OH co-product. This dissociation pathway has
never been reported in ester before. Despite a large available energy above the bond

13
fission of CH3O-CHO in radical dissociation channel, a small portion of CH3O has a
chance to undergo intramolecular abstraction at some distance. In this manner, the
resulting available energy deposition in CH3OH and CO shows distinct difference
from those via the conventional transition state.

Acknowledgments

This work is supported by National Science Council of Taiwan, Republic of


China under contract no. NSC 99-2113-M-001-025-MY3. Computer resources at the
National Center for High-performance Computer were utilized in the calculations.

14
References

1. S. M. Japar, T. J. Wallington, J. F. O. Richert and J. C. Ball, Int. J. Chem.


Kinet., 1990, 22, 1257.
2. M. E. Jenkin, G. D. Hayman, T. J. Wallington, M. D. Harley, J. C. Ball, O. J.
Nielsen and T. Ellerman, J. Phys. Chem., 1993, 73, 11712.
3. J. Wenger, E. Porter, E. Collins, J. Treacy and H. Sidebottom, Chemisphere,
1999, 38, 1197.
4. M. A. Barteau and R. J.Madix, J. Catal., 1980, 62, 329.
5. T. Ushikubo, H. Hattori and K. Tanabe, Chem. Lett., 1984, 13, 649.
6. F. Q. Ma, D. S. Lu and Z. Y. Guo, J. Catal., 1992, 134, 644.
7. F. Q. Ma, D. S. Lu and Z. Y. Guo, J. Mol. Catal., 1993, 78, 309.
8. E. W. R. Steacie, Proc. R. Soc. (London), 1930, 4127, 314.
9. D. V. S. Jain and B. S. Murwaha, Indian J. Chem., 1969, 7, 901.
10. C. P. Davis, Ph.D. Thesis, University of Mississippi, Oxford, Mississippi, 1983.
11. J. S. Francisco, J. Am. Chem. Soc., 2003, 125, 10475.
12. R. C. L. Pereira and P. C. Isolani, J. Photochem. Photobiol., A, 1988, 42, 51.
13. S.H. Lee, J. Chem. Phys., 2008, 129, 194304.
14. J. G. Lo´pez, G. Vayner, U. Lourderaj, S. V. Addepalli, S. Kato, W. A. deJong,
T. L. Windus and W. L. Hase, J. Am. Chem. Soc., 2007, 129, 9976.
15. A. E. Pomerantz, J. P. Camden, A. S. Chiou, F. Ausfelder, N. Chawla, W. L.
Hase and R. N. Zare, J. Am. Chem. Soc., 2005, 127, 16368.
16. L. Sun, K. Song and W. L. Hase, Science, 2002, 296, 875
17. U. Lourderaj and W. L. Hase, J. Phys. Chem. A, 2009, 113, 2236
18. P. L. Houston and S. H. Kable, Proc. Natl. Acad. Sci. U.S.A., 2006, 103, 16079.
19. 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, 2004, 306,
1158.
20. A. G. Suits, Acc. Chem. Res., 2008, 41, 873.
21. L.B. Harding, S.J. Klippenstein and A.W. Jasper, Phys. Chem. Chem. Phys.,
2007, 31, 4055.

15
22. H. M. Yin, S. H. Kable, X. Zhang and J. M. Bowman, Science, 2006, 311, 1443.
23. X. Zhang, J. L. Rheinecker and J. M. Bowman, J. Chem. Phys., 2005, 122,
114313.
24. K. M. Christoffel and J. M. Bowman, J. Phys. Chem. A, 2009, 113, 4138.
25. V. Goncharov, S. A. Lahankar, J. D. Farnum, J. M. Bowman and A.G. Suits, J.
Phys. Chem. A, 2009, 113, 15315.
26. L. B. Harding, Y. Georgievskii and S. J. Klippenstein,. Proc. Natl. Acad. Sci.
U.S.A., 2008, 105, 12719.
27. L. Rubio-Lago, G. A. Amaral, A. Arregui, J. G. Izquierdo, F. Wang, D.
Zaouris, T. N. Kitsopoulos and L. Banares, Phys. Chem. Chem. Phys., 2007, 9,
6123.
28. R. Sivaramakrishnan, J. V. Michael and S. J. Klippenstein, J. Phys. Chem. A,
2010, 114, 755.
29. B. C. Shepler, B. J. Braams and J. M. Bowman, J. Phys. Chem. A, 2008, 112,
9344.
30. L. B. Harding, Y. Georgievskii and S. J. Klippenstein, J. Phys. Chem. A, 2010,
114, 765.
31. S. Chen and W. H. Fang, J. Chem. Phys., 2009, 131, 054306.
32. V. Goncharov, N. Herath and A. G. Suits, J. Phys. Chem. A, 2008, 112, 9423.
33. S. Saxena, J. H. Kiefer and S. J. Klippenstein, Proceedings of the Combustion
Institute, 2009, 32, 123.
34. S. Maeda, K. Ohno and K. Morokuma, J. Phys. Chem. Lett., 2010, 1, 1841.
35. M. P. Grubb, M. L. Warter, A. G. Suits and S. W. North, J. Phys. Chem. Lett.,
2010, 1, 2455.
36. C. Chen, B. Braams, D. Y. Lee, J. M. Bowman, P. L. Houston and D. Stranges,
J. Phys. Chem. Lett., 2010, 1, 1875.
37. A. T. J. B. Eppink and D. H. Parker, Rev. Sci. Instrum., 1997, 68, 3477.
38. D. H. Parker and A. T. J. B. Eppink, J. Chem. Phys., 1997, 107, 2357.
39. Y. Tang, W. B. Lee, Z. F. Hu, B. Zhang and K. C. Lin, J. Chem. Phys., 2007, 126,
064302.
40. X. P. Zhang, Z. R. Wei, Y. Tang, T. J. Chao, B. Zhang and K. C. Lin,

16
ChemPhysChem, 2008, 9, 1130.
41. X. P. Zhang, W. B. Lee, D. F. Zhao, M. K. Hsiao, Y. L. Chen and K. C. Lin, J.
Chem. Phys., 2009, 130, 214305.
42. M. Castillejo, S. Couris, E. Lane, M. Martin and J. Ruiz, Chem. Phys. 1998, 232,
353.
43. V. Dribinski, A. Ossadtchi, V. A. Mandelshtam and H. Reisler, Rev. Sci.
Instrum. 2002, 73, 2634.
44. E. E. Barnes and W. T. Simpson, J. Chem. Phys., 1963, 39, 670.
45. E. Vésine and A. Mellouki, J. Chim. Phys. Phys. Chim. Biol, 1997, 94, 1634.
46. A. B. Rocha, A. S. Pimentel and C. E. Bielschowsky, J. Phys. Chem. A, 2002,
106, 181.
47. G. Herzberg and D. A. Ramsay, Proc. R. Soc. London, Ser. A, 1955, 233, 34.
48. J. S. Francisco, A. N. Goldstein and I. H. Williams, J. Chem. Phys., 1988, 89,
3044.
49. L. R. McCunn, M. J. Krisch, K. Takematsu, L. J. Butler and J. N. Shu, J. Phys.
Chem. A, 2004, 108, 7889.
50. Z. R. Wei, X. P. Zhang, W. B. Lee, B. Zhang and K. C. Lin, J. Chem. Phys.,
2009, 130, 014307.
51. NIST web page: http://srdata.nist.gov/cccbdb/.
52. D. A. Good, J. Hanson, M. Kamoboures, R. Santiono and J. S. Francisco, J.
Phys. Chem. A, 2000, 104, 1505.

17
Table 1. Energy deposition and branching ratio of bimodal translational energy
distribution of CO as a function of rotational level.

Roaming
J”CO Sharp band (kcal/mol) Broad band (kcal/mol)
ratio
<ET> a <Eint, CH3OH >b <ET> <Eint, CH3OH >

23.4 83.4
24 ~ 0%
(20.1%) (73.6%)
2.3 105.7 23.9 84.1
19 2.7±1.1%
(2.0%)c (93.4%) (21.1%) (74.3%)
2.1 106.8 23.3 85.7
14 5.5±2.1%
(1.9%) (94.4%) (20.5%) (75.7%)
1.4 108.1 22.6 86.9
10 7.3±0.6%
(1.1%) (95.5%) (19.9%) (76.8%)

a. <ET> denotes the average translational energy.


b. <Eint, CH3OH > is the internal energy deposition in CH3 OH, evaluated by the
formula: <Eint, CH3OH > = Eavl - <ET> - <Eint, CO>. Internal energy of CO(v”=0,
J”) is estimated by summing the zero-point energy of 1081.59 cm-1 and
individual rotational energy of 211.4, 403.5, 729.7, and 1151.3 cm-1
corresponding to J”=10, 14, 19, and 24, respectively.
c. The data in parenthesis is a ratio shared by each energy partition. The ratio
of <Eint, CO > is given by (100%- <ET>- <Eint, CO >).

18
Figure Captions

Fig.1 Rotational level dependence of (a) raw ion images and (b) corresponding
center-of-mass translational energy distributions of the CO fragment in the
photolysis of methyl formate at 234 nm. The translational energy distribution
is deconvoluted into one broad and one sharp component according to eq.3.

Fig.2 Power dependence of CO(J”=10) in the photodissociation of methyl formate at


234 nm, yielding a slope of 1.20.1.

Fig.3 Unimolecular decomposition pathways of methyl formate in units of kcal/mol.


The results are calculated using B3LYP/cc-pVTZ//CCSD(T)/cc-pVTZ level of
theory, in consistency with those data in parenthesis calculated by Francisco.
The enthalpy of reaction for the atomic channel of HCOOCH3 CH3OCO +
H is adopted from ref.52.

19
(a) (b)

Fig.1

20
Fig.2

21
Fig.3

22

You might also like