Вы находитесь на странице: 1из 15

PCCP

View Article Online


PERSPECTIVE View Journal | View Issue

Molecular halogen elimination from


Published on 21 February 2014. Downloaded by Academia Sinica - Taipei on 15/12/2017 09:16:31.

halogen-containing compounds in the atmosphere


Cite this: Phys. Chem. Chem. Phys.,
2014, 16, 7184
King-Chuen Lin* and Po-Yu Tsai

Atmospheric halogen chemistry has drawn much attention, because the halogen atom (X) playing a
catalytic role may cause severe stratospheric ozone depletion. Atomic X elimination from X-containing
hydrocarbons is recognized as the major primary dissociation process upon UV-light irradiation, whereas
direct elimination of the X2 product has been seldom discussed or remained a controversial issue. This
account is intended to review the detection of X2 primary products using cavity ring-down absorption
spectroscopy in the photolysis at 248 nm of a variety of X-containing compounds, focusing on
bromomethanes (CH2Br2, CF2Br2, CHBr2Cl, and CHBr3), dibromoethanes (1,1-C2H4Br2 and 1,2-C2H4Br2) and
dibromoethylenes (1,1-C2H2Br2 and 1,2-C2H2Br2), diiodomethane (CH2I2), thionyl chloride (SOCl2), and
sulfuryl chloride (SO2Cl2), along with a brief discussion on acyl bromides (BrCOCOBr and CH2BrCOBr).
Received 15th November 2013, The optical spectra, quantum yields, and vibrational population distributions of the X2 fragments have
Accepted 18th February 2014 been characterized, especially for Br2 and I2. With the aid of ab initio calculations of potential energies
DOI: 10.1039/c3cp54828g and rate constants, the detailed photodissociation mechanisms may be comprehended. Such studies are
fundamentally important to gain insight into the dissociation dynamics and may also practically help to
www.rsc.org/pccp assess the halogen-related environmental variation.

1. Introduction these source gases to the stratosphere and thus reduce the
amount of reactive free radicals formed. Regarding their tropo-
Atmospheric halogen chemistry has drawn much attention, spheric lifetimes, for instance, trihalomethanes including
because the emission of halocarbons into the troposphere CHBr2Cl, CHBrCl2, and CHBr3 were reported to last for about
may potentially cause severe damage to the stratospheric ozone 120 days by the reaction with OH, whereas their lifetimes with
layer. There are several catalytic cycles involving reactive free respect to photolysis near the earth’s surface (0–5 km) were
radicals of Cl, ClO, Br, and BrO that cause chlorine and found to be 20–60 days.6 The way to remove these pollutants is
bromine to destroy effectively stratospheric ozone.1–4 Iodine mainly through the UV photodissociation processes. Therefore,
chemically coupled with emissions of bromine and chlorine is understanding the related UV photochemistry is extremely
expected to account for the widespread depletion of lower important for the assessment of their environmental impact.
stratospheric ozone below 20 km altitude.5 A short tropospheric In the halogen photochemistry, atomic halogen (X) elimination
lifetime of halocarbons can essentially prevent transport of is recognized as the major primary dissociation process,
whereas direct elimination of molecular halogen (X2) has been
Department of Chemistry, National Taiwan University, Taipei, and Institute of
seldom discussed. By calculating the ozone loss rate either per
Atomic and Molecular Sciences, Academia Sinica, Taipei 106, Taiwan. inorganic bromine or chlorine atom, Daniel et al. suggested
E-mail: kclin@ntu.edu.tw; Fax: +886-2-23621483 that the bromine is about 45 times more effective than chlorine

King-Chuen Lin is a distinguished professor of the Department of Po-Yu Tsai is a postdoctoral fellow of the Department of Chemistry
Chemistry at National Taiwan University and a distinguished at National Taiwan University. He received his BS degree in
research fellow of National Science Council, Taiwan. He received Chemistry from Tunghai University and his PhD in Chemistry from
his BS degree in Chemistry from National Taiwan University, National Taiwan University, Taiwan. His research focuses on
Taiwan, his PhD in Chemistry from Michigan State University, energy transfer reactions in the gas phase, photodissociation
USA, and his postdoctoral degree from Cornell University. His dynamics of oriented molecules, and atmospheric chemistry.
research interests are photodissociation and reaction dynamics in
gas and condensed phases, atmospheric chemistry, and single
molecule spectroscopy.

7184 | Phys. Chem. Chem. Phys., 2014, 16, 7184--7198 This journal is © the Owner Societies 2014
View Article Online

Perspective PCCP

in global ozone destruction.7 Photochemistry of bromine-containing an important impact on fundamental understanding of the
hydrocarbons has received much less attention compared to that of dissociation dynamics and practical help with the photo-
chlorine-containing hydrocarbons. Thus, this account is intended to chemical assessment of halogen-related environmental issues.
stress on detection, characterization, and dissociation mechanisms This account is intended to review the CRDS detection of the
associated with the Br2 primary elimination channel, but discuss Br2, I2, and Cl2 primary products in the photolysis at 248 nm of
briefly the Cl2 and I2 elimination processes. a variety of halogen-containing compounds, focusing mainly on
It has been a controversial issue whether the X-containing bromomethanes (CH2Br2, CF2Br2, CHBr2Cl, and CHBr3), dibromo-
hydrocarbons may release the X2 primary products upon UV-light ethanes (1,1-C2H4Br2 and 1,2-C2H4Br2), dibromoethylenes
irradiation. Since the early 1960s, the CF2 product has been (1,1-C2H2Br2 and 1,2-C2H2Br2), iodomethane (CH2I2), thionyl
Published on 21 February 2014. Downloaded by Academia Sinica - Taipei on 15/12/2017 09:16:31.

observed in the photolysis of C2F2Br2 at 248 nm, and was ascribed chloride (SOCl2), and sulfuryl chloride (SO2Cl2), along with a
to a process of the Br2 elimination.8,9 But the Br2 moiety has not brief discussion on acyl bromides (BrCOCOBr and CH2BrCOBr).
been detected directly. In contrast, the same molecule at the The optical spectra, quantum yields, and vibrational population
photolysis wavelengths of 223–260 nm was studied using a distributions of the molecular halogen elimination will be
crossed laser-molecular beam technique, suggesting that (1) a characterized. With the aid of calculations of ab initio reaction
single C–Br bond fission is the only primary reaction channel, pathways and rate constants, we can gain insight into the
and (2) CF2 should be accompanied by two Br fragments.10,11 detailed photodissociation mechanisms.
The photodissociation of SOCl2 is another example. The Cl2 + SO
channel was verified following irradiation at 248 nm by observa-
tion of a mass/charge ratio of 70 using photofragment transla- 2. Experimental
tional spectroscopy (PTS).12 In contrast, a recent investigation
found no evidence of such a channel (SOCl2 - SO + Cl2) using The CRDS apparatus used for the photodissociation study is
three-dimensional imaging of photofragments performed described elsewhere.29–31 The radiation sources are composed
at 235 nm.13 of a 20 ns-pulsed excimer laser emitting at 248 nm for photo-
Even in the photolysis of halogen-substituted ethylenes, the lysis of precursors and a 5–8 ns-pulsed Nd:YAG laser-pumped
molecular halogen elimination has been rarely characterized.14–20 dye laser (482–524 nm) used to probe the released X2 fragment
For instance, Sato et al. have observed HCl, Cl, C2H2, C2HCl, and in the B3Pou+ ’ X1Sg+ transition.29–33 The energies of photo-
C2Cl2 fragments in dichloroethylenes and trichloroethylenes at lysis and probe lasers were controlled in the range of 8–28 and
193 and 157 nm.14 He et al.18 and Chandra et al.19 have detected 0.5–1 mJ, respectively. The photolysis laser beam was focused
H, HCl, Cl(2P3/2), and Cl*(2P1/2) in 1,1- and 1,2-C2H2Cl2 at 193 nm with a 30 cm focal-length cylindrical lens onto a four-armed
and other UV wavelengths. But the prior studies have never found stainless steel ring-down cell at right angle to the cavity, while
the Cl2 fragments. From a theoretical point of view, Morokuma the probe laser beam was injected along the cavity axis after a
and coworkers20 suggested that a high energy barrier might 20–30 ns delay (Fig. 1). The two beams were overlapped in the
impede the release of the Cl2 product. In 1,2-dibromoethylene center of the flow cell. The volume of the overlapping region
at 248 nm by PTS,21 Lin and coworkers claimed to have found two was evaluated by multiplication of the beam width and height
exclusive dissociation channels of Br2 + C2H2 and Br(fast) + of the photolysis laser and the beam diameter of the probe
Br(slow) + C2H2 with a branching ratio of B0.2 : 0.8. However, laser, corresponding to (18  1  3)  5 mm3. In order to
such a Br2 channel could not be observed using velocity ion remain mostly in the TEM00 mode, the probe beam was guided
imaging by our group,22 nor by the Lipson group over a through a spatial filter made of a pair of lenses of 10 cm and
different range of laser wavelengths.23 5 cm focal length and a pinhole with 70 mm diameter, prior
It is difficult to detect the neutral X2 by using the resonance- to injection through the front mirror of the ring-down cell.
enhanced multiphoton ionization (REMPI) technique. For
instance, the REMPI detection of the Br2 signal suffers from
(1) competition of Br2+ interference resulting from dissociative
ionization,24 e.g., RBr2 + 2hv - RBr2+ + hv - R + Br2+, where
R denotes a hydrocarbon, and (2) partial Br2 dissociation
that may be caused concomitantly by the UV light applied.25
Even using a laser-induced fluorescence (LIF) technique, it is
inefficient to probe Br2 in the B3Pou+ ’ X1Sg+ transition, due to
a low fluorescence quantum yield caused by predissociation
with a repulsive C1P1u state.26 In contrast, as an emerging
absorption technique, cavity ring-down absorption spectro-
scopy (CRDS) turns out to be superior to most spectroscopic
techniques in detecting the halogen molecules.27
We have recently taken advantage of the CRDS method28 to
probe the X2 (Cl2, Br2, and I2) fragments in a series of halo- Fig. 1 Schematic diagram of the stainless steel reaction chamber adopted
alkanes, haloalkenes, and acyl halides. Such investigation has for cavity ring-down absorption spectroscopy.

This journal is © the Owner Societies 2014 Phys. Chem. Chem. Phys., 2014, 16, 7184--7198 | 7185
View Article Online

PCCP Perspective

A photomultiplier tube was positioned behind the rear mirror to


detect the intensity of the light pulse leaking out of the mirror. The
temporal profile of the ring-down signal was recorded on a transient
digitizer. The ring-down time constant for each laser pulse was
determined by a best fit to the acquired exponential decay.
The precursors were purified by repeated freeze–pump–thaw
cycles at 77 K and then flowed through the ring-down cell at a
pressure of 10–500 mTorr monitored by an MKS pressure
gauge. The X2 absorption spectra, with a spectral resolution
Published on 21 February 2014. Downloaded by Academia Sinica - Taipei on 15/12/2017 09:16:31.

of 0.1 cm1, were acquired with the aid of a lab-developed


program based on a Matlab environment. For the temperature-
dependence measurements, the whole chamber was wrapped
with a heating tape and the varied temperature was monitored
by a thermocouple positioned nearby the center region.
Fig. 2 A portion of Br2 spectra acquired in the photolysis of BrCOCOBr at
248 nm. (a) Trace acquired experimentally for the bands of v = 0, 1, and
2 contained in the 534–535 nm region, (b) the simulated counterpart
3. Results and discussion with the population ratio of Br2(v = 1)/Br2(v = 0) fixed at 0.65 and Br2(v = 2)/
Br2(v = 0) optimized to 0.34, (c) simulated counterpart with the transition
3.1. Photodissociation of Br-substituted methanes involving only v = 0, (d) simulated counterpart with the transition involving only
v = 1, and (e) simulated counterpart with the transition involving only v = 2.
3.1.1. Br2 production. CH2Br2 is one of the major bromine-
containing organic compounds, contributing up to 20% of Br
atoms into the upper troposphere upon UV-light irradiation.34 while excitation at 210 nm within the B band leads predomi-
But the work on its photodissociation dynamics is limited.35–37 nantly to the C–Br bond cleavage.41,42
CF2Br2 (Halon 1202) is anticipated to yield three major The CRDS spectra of Br2 are acquired at room temperature,29–31
dissociation channels of CF2Br + Br, CF2 + 2Br, and CF2 + Br2 containing the Br2 v = 0 transition from the bands (35,0) to (44,0) in
that are energetically attainable in the ultraviolet photolysis.11 the 515.5–518 nm range, and both v = 0 and 1 transitions from
But it has not been confirmed whether a primary dissociation (30,0) to (33,0) and from (36,1) to (44,1) in the 522–524 nm
pathway can produce CF2 + Br2. By using femtosecond pump– range. The Br2 signals, free from interference of other species,
probe spectroscopy, Dantus and coworkers investigated the disappear when the photolysis laser is off. The CRDS spectra for
UV-multiphoton dissociation of several dihalomethanes,36,37 v = 2 were not observed for these molecules, but acquired for
obtaining the halogen products lying in the excited states. BrCOCOBr in the 534–535 nm region, as shown in Fig. 2.31
The related mechanisms essentially differ from those found The spectral assignment is made according to the report
in the nanosecond domain. Bromoform is another long-lived by Barrow et al.43 The Br2 (v = 0, 1, and 2) spectral intensities
source of atmospheric bromine, the UV photodissociation of were simulated27,31 and the results for the vibrational popula-
which results in the fragments of HBr, CBr, CHBr, CHBr2, Br, tion ratios are listed in Table 1. The Br2 fragment obtained in
and Br2.38 North and coworkers38 using the PTS technique each molecule is essentially vibrationally hot, according to the
demonstrated that the Br2 product originates from a secondary Boltzmann law.
dissociation of CHBr2, whereas Xu et al. using a velocity ion The details of spectral simulation are described in the following.
imaging technique claimed the observation of the Br2 primary The rotational and vibrational constants of the 79Br2 isotopic
channel in the 234 and 267 nm photolysis.39 But they measured variant were determined by Gerstenkorn and Luc.44 The molecular
indirectly the CHBr and CHBr2 intensities that were used to constants of 79,81Br2 and 81Br2 are then evaluated based on the
determine the branching ratios of Br2 and Br elimination. This isotope ratio. Given these molecular constants of the X1Sg+
account reports the first finding of the Br2 primary product in and B3Pou+ states for all the isotopic variants, the Br2 CRDS
the CRDS experiments. For comparison, another three-halogen- spectra may be assigned accurately. The Br2 (v = 0) spectral
substituted methane CHBr2Cl was selected for the CRDS
experiments. Since the bond dissociation energy of C–Cl is
302 kJ mol1 forming CHBr2 + Cl with respect to the C–Br Table 1 Quantum yields and vibrational population ratios of various
dissociation energy of 247 kJ mol1 into CHBrCl + Br,29,40 the multi-Br-substituted hydrocarbons
Cl-atom loss should be significantly less than the Br-atom loss
Quantum yield v = 1/v = 0 ratio
at 248 nm with the A band excited and thus the interference
from the secondary photodissociation of CHBr2 is minimized CHBr3 0.23  0.05 0.8  0.2
CF2Br2 0.04  0.01 0.4  0.2
while detecting the Br2 molecular product. However, it should CH2Br2 0.2  0.1 0.7  0.2
be noted that when the B band is excited, the efficiency of CHBr2Cl 0.05  0.03 0.5  0.2
Cl- and Br-atom loss might be inverted. For instance, for 1,2-C2H2Br2 0.12  0.1 0.7  0.2
1,1-C2H2Br2 0.07  0.04 0.55  0.05
photodissociation of CH2BrI, excitation at 248 nm within the 1,2-C2H4Br2 0.36  0.18 0.8  0.1
A absorption band preferentially leads to the C–I bond cleavage, 1,1-C2H4Br2 0.05  0.03 0.5  0.2

7186 | Phys. Chem. Chem. Phys., 2014, 16, 7184--7198 This journal is © the Owner Societies 2014
View Article Online

Perspective PCCP

intensities I can be further simulated according to the following addition to vibrational population ratios. The Br2 yield in
equation: CHBr3 equals 0.23  0.05 which is overestimated,48 containing
partial contribution from the secondary dissociation of CHBr2.38
ðFCFÞðHLFÞ
I ¼k NJ 00 riso f (1) In comparison, the quantum yield in CHBr2Cl is 0.05  0.03.29
2J 00 þ 1
While considering the structural analog, this value may well be
where k is a factor associated with instrument and experimental taken as the lower limit of the Br2 yield in CHBr3. Between these
conditions, FCF the Franck–Condon factor,45 HLF the Hönl– two compounds, a higher photolyzing laser energy is required
London factor, J 00 the rotational quantum number of the ground to decompose CHBr3, implying that a large fraction of the
state Br2, NJ 00 the Boltzmann distribution of the rotational Br2 product comes from multiphoton absorption processes. In
Published on 21 February 2014. Downloaded by Academia Sinica - Taipei on 15/12/2017 09:16:31.

population for which 300 K is assumed, because the rotational CF2Br2, the quantum yield is 0.04  0.01. Troe and coworkers
population has been thermally equilibrated during a long ring- have used the IR multiphoton absorption technique to probe the
down time. riso the ratio of isotopic variants, 79Br2 : 79,81Br2 : 81Br2 Br2 channel with a branching ratio of 0.10–0.25, depending on
equal to 0.2569 : 0.4999 : 0.2431, and f the intensity ratio of strong the IR laser energy.49 This range may be considered as an upper
and weak rotational lines. Because the nuclear spin of Br is 3/2, limit of the Br2 yield at 248 nm.
f becomes 5/3 for the intensity ratio of odd to even rotational 3.1.2. Verification of the primary product of Br2. To inspect
quantum number.46 Given BrCOCOBr as an example, while the possibility of atomic Br recombination during the acquisi-
taking into account all the band transitions involved with the tion period with a ring-down time o1 ms, a single-Br-containing
same line width of 0.1 cm1 and the related spectral intensities substitute CCl3Br or CH2ClBr with a comparable absorption
by eqn (1), a spectral simulation for the Br2 (v = 0) population is cross section was photolyzed.50 None of the Br2 fragments can
obtained in the 515.5–518 nm range. Then, a theoretical be detected, even with increased pressure and prolonged photolysis-
counterpart is analogously obtained in the 522–524 nm range probe delay time (Fig. 3). To verify the primary product of Br2, two
by adjusting the Br2(v = 1)/Br2(v = 0) population ratio. As the further experiments involving laser power and pressure dependence
Br2(v = 1)/Br2(v = 0) ratio is fixed, a simulated spectrum in the were performed. A straight line is obtained in a plot of the rotational
534–535 nm region is optimized by adjusting the population intensity versus the incident laser energy, indicative of a single-
ratio of Br2(v = 2)/Br2(v = 0). In this manner, the theoretical photon involvement in the molecular elimination. Each plot for
counterpart in Fig. 2 is obtained, showing consistency with the CH2Br2, CF2Br2, and CHBr2Cl has taken into account the origin
experimental findings. point in the linear regression fit, thereby ruling out the effect of
The quantum yield of the Br2 channel can be determined optical saturation or partial saturation. The related plots for
as follows: CH2Br2 as an example are shown in Fig. 4. However, for CHBr3
½Br2  the line cannot be extrapolated through the origin; a partial
f¼ ; (2) contribution from the CHBr2 secondary dissociation should be
Np
significant, as described above.38 The pressure-dependence
where [Br2] indicates the Br2 concentration produced in the measurements also yield a straight line, indicative of the Br2
beam-crossed region of photolysis/probe lasers and Np is the production from a single molecular photodissociation. The
photon number density absorbed in the same region. The Br2 linearly dynamic range of the pressure dependence differs for
concentration is evaluated by a ratio of a/sBr2, in which a is the each molecule studied. When the pressure keeps increasing,
absorption coefficient and sBr2 the absorption cross section.27,47 the Br2 product intensity becomes level-off and rapidly tends to
a is determined by the ring-down time measurements, decrease. The complicated high pressure phenomena might be
  caused by quenching of the excited state population of the
d 1 1
a¼  ; (3)
cl t t0

where d is the distance between two reflective mirrors; l the


optical length of the absorber; c the light speed, and t and t0
are the ring-down times of Br2 as a result of the probe
laser wavelength in-resonance and off-resonance, respectively.
Meanwhile, the absorbed photon number density can be evaluated
by using the following equation,
Ein  Eout
Np ¼ ; (4)
hvDV
where h denotes Planck’s constant, n is the radiation frequency,
DV is the volume of the beam-crossed region, and Ein and Eout
indicate the beam intensity of incoming and outgoing radiation
Fig. 3 Detection of CRDS spectra of Br2 following 248 nm photolysis of
for photolysis. (a) CCl3Br with a pressure of 2.2 Torr and the photolysis-probe delay
Accordingly, the Br2 quantum yields for the molecules time of 80 ns and (b) CF2Br2 with a pressure of 0.5 Torr and a delay time
studied are evaluated and the results are listed in Table 1, in of 20 ns.

This journal is © the Owner Societies 2014 Phys. Chem. Chem. Phys., 2014, 16, 7184--7198 | 7187
View Article Online

PCCP Perspective

diethyl ether acting as the Br scavenger has a larger reaction


rate constant than with BrCOCOBr. If the added amount of
diethyl ether exceeds the precursor, then the generated Br
fragment must be largely consumed with little chance to proceed
via the Br + precursor reaction. The resulting Br2 spectra do not
show any suppression by the Br scavenger, thus indicative of an
insignificant Br2 contribution from such a secondary reaction.
However, it is not easy to find an appropriate Br scavenger,
because it may concomitantly quench the excited precursor
Published on 21 February 2014. Downloaded by Academia Sinica - Taipei on 15/12/2017 09:16:31.

rapidly to reduce the efficiency of Br2 elimination.27


There is one more possibility to generate Br2 through dissocia-
tion of adducts which may be formed between halogen atoms and
haloalkanes. Enami et al.54 have claimed the observation of
adducts in the reaction of Cl with XCH2I (X = H, CH3, Br, Cl,
and I) in 25–125 Torr of N2 diluent at 250 K. Sehested et al.55 also
found two reaction pathways in the reaction of F with CH3Br,
yielding (1) CH2Br and HF, and (2) the adduct CH3Br  F. The F
atoms were produced using pulse radiolysis of 1 atm of SF6. If the
adducts further undergo secondary decomposition, the halogen
molecule is possibly produced. However, under our experimental
conditions, Br2 could not be generated via dissociation of such
adducts. First, the adduct formation requires a high pressure
matrix added for stabilization by termolecular collisions.54,55
Second, the subsequent elimination of Br2 has a very small
rate, because of endothermicity. Third, if the adducts require
one additional photon to release Br2, two-photon absorption is
necessary to form one Br2 molecule that is against our observa-
tion of laser energy dependence.
3.1.3. Photodissociation mechanisms. How does the Br2
elimination occur in the photolysis of these bromine-
containing molecules? It has been well recognized that the first
Fig. 4 (a) Line intensity of Br2 rotation at 519.68 nm eliminated from
UV-absorption band (n, s*) of bromo-alkanes at 248 nm readily
CH2Br2 versus the incident laser energy. The origin point is taken into dissociates a single Br atom.51,52 It is hard to eliminate a Br2
account for the linear regression fit. The same plot with the corresponding molecule from this excited state. For instance, the excited state
number density of Br2 against the absorbed photon density yields a slope, of CH2Br2 promoted at 248 nm has a 11,3B1 symmetry.56 Such
indicative of the quantum yield of 0.2  0.1 for the channel of molecular
excited molecules are symmetry-forbidden to eliminate Br2,
elimination. (b) Line intensity of Br2 rotation at 519.68 nm versus the
CH2Br2 pressure from 0.7 to 3.1 Torr.
without undergoing nonadiabatic transition to the ground
potential surface. As the C–Br bond is elongated, the CH2Br2
structure changes from a C2v to Cs symmetry and the subse-
precursor prior to photodissociation, quenching of the quent 11A 0 –X̃1A 0 coupling becomes strengthened to facilitate
ro-vibrational population of the produced Br2, and the possibility the internal conversion (IC). A small fraction of population may
of secondary reactions. have a chance to transit to the higher vibrational levels of
The Br2 fragment may be alternatively produced by the the ground state X̃1A 0 prior to the C–Br bond fission. The
dissociated Br atoms in reactions with the precursor. A single experiments of the temperature effect on Br2 elimination may
atomic halogen elimination has been well known to dominate support the above photodissociation mechanism. For the mole-
the UV-decomposition channels of the precursors studied.51–53 cules studied, the intensities were enhanced from 5% to 16%
Thus, the resultant Br yields should be so large that the pseudo- within the temperature increment from 20 to 50 1C.29,56,57
first order reaction cannot be applied to the above reactions. In As the temperature increases, higher levels of excited states
this sense, the Br2 production requires two-precursor involve- are populated such that the increased density of states may
ment, which is against the linear pressure-dependence mea- facilitate the total rate of the level-to-level coupling, resulting in
surements. Furthermore, such a reaction is endothermic with a a more efficient IC process.
slow reaction rate, because a large C–Br dissociation energy is The detailed photodissociation mechanisms can be under-
accompanied by a small Br2 formation energy in the bond stood with the aid of potential energy surface (PES) calcula-
breaking/formation process. Alternatively, a further experiment tions. The Br2 and probable dissociation channels on the
may be carried out by adding a Br scavenger in the ring-down adiabatic singlet ground-state (or first triplet state for some
cell. Given BrCOCOBr as an example,31 the Br reaction with cases) PES for each compound are characterized. The geometries

7188 | Phys. Chem. Chem. Phys., 2014, 16, 7184--7198 This journal is © the Owner Societies 2014
View Article Online

Perspective PCCP

and the harmonic frequencies of reactants, transition states Note that the isomerization photoproducts detected in the
(TS), and products are obtained at the level of the hybrid condensed phase have a mechanism different from the isomer-
density functional theory, B3LYP/6-311G(d,p); the energies are ization TS along the dissociation pathway in the gas phase.61
further refined by the coupled cluster CCSD(T)/cc-pVTZ with
B3LYP/6-311G(d,p) zero-point energy corrections. The vertical 3.2. Photodissociation of Br-substituted ethanes
energies of the lowest excited states with respect to the ground The isomeric forms of 1,1- and 1,2-dibromoethanes are photo-
state are determined by the Davidson-corrected multi-reference lyzed at 248 nm, both yielding the Br2 fragments on the ground
configuration interactions with single and double excitation state PESs. Nevertheless, their photodissociation processes
(MRDCI) calculations, in which the cc-pVTZ basis functions are differ distinctly from each other.30 The Br2 yield and branching
Published on 21 February 2014. Downloaded by Academia Sinica - Taipei on 15/12/2017 09:16:31.

employed. The core electron correlation is not involved, because ratio of Br2(v = 1)/Br2(v = 0) in 1,2-dibromoethane is 0.36 and
of its minor effect on the relative energy.58,59 1,1-dibromoethylene 0.8  0.1 in contrast to 0.05 and 0.5  0.2 in 1,1-dibromoethane.
is examined herein via a calculation level of full-electron CCSD(T)/ In contrast, the Br2 number density increases by a factor of
Sapporo-DZP-201260 showing a relative energy deviation less than 35% versus 190%, when the temperature is increased from
1.3 kcal mol1 with respect to a frozen core approximation, 34 to 52 1C.
whereas the consumed CPU time is four times more. For simpli- The discrepancy can be interpreted with the aid of the PES
city, the spin–orbit splitting of the Br atom is not taken into calculations. As shown in Fig. 5 and 6, the TS energies lie at
account for the ab initio calculations, but this would not have any 334.1 and 492.4 kJ mol1 for 1,2- and 1,1-dibromoethanes,
effect on the mechanism proposed for the Br2 elimination. respectively. The former state is well below the excitation
In the photodissociation of CHBr2Cl,29 the TS correlates to energy at 483 kJ mol1 (248 nm), whereas the latter is slightly
the CHCl + Br2 products, following a reaction coordinate of the above. Their excited density of states and the IC efficiency
ground state PES. For the optimized TS structure, the C–Br are approximately equal. The production rates are thus deter-
bond lengths are elongated asymmetrically and the Br–Br distance mined mainly by the transition barrier. Because a small energy
is largely stretched. Thus, a mechanism of asynchronous photo- barrier impedes the photodissociation of the ground state
dissociation is favored and the Br2 population is vibrationally hot. 1,1-dibromoethane, the Br2 yield becomes relatively low, but
CH2Br2 follows a similar mechanism to produce vibrationally hot rises rapidly with the temperature according to Arrhenius
Br2 fragments.56 In contrast, a sequential photodissociation theory. In 1,2-dibromoethane, Br2 is anticipated to eliminate
mechanism is favored in CF2Br2.57 That is, a single C–Br bond via a four-center mechanism with the Br–Br bond distance of
breaks first, and then the free Br atom moves to form a Br–Br 2.904 Å in the optimized TS structure, whereas in 1,1-dibromoethane
bond in the TS, followed by the Br2 elimination. Reid and Br2 is via a three-center elimination with the bond distance of
coworkers61 have extensively calculated the dissociation path- 2.543 Å. The difference of the Br–Br stretch bond might explain
ways for multi-halocarbons including CF2Cl2, CF2Br2, and why the former isomer leads to a hotter Br2.
CHBr3, and demonstrated that the molecular products are
dominated by the pathway through such an isomerization TS 3.3. Photodissociation of Br-substituted ethylenes
carrying a halogen–halogen bond. For instance, the decompo- The prior studies of photodissociation mechanisms of halo-
sition of CF2Cl2 via an isomerization TS CF2Cl  Cl gives rise to ethylenes,66–69 denoted as C2H4–nXn, yield at least three dissocia-
a 3% Cl2 yield in appreciable agreement with the findings tion channels. They are (1) four-center elimination to release
in infrared multiphoton dissociation reported by Lee and HX, H2, HCCH or HCCX, (2) three-center elimination to release
coworkers.62 In CHBr3, the three-center TS reported previously48 HX, H2, HCCH or HCCX via vinylidene or halovinylidene
lies at B139 kJ mol1 higher than the calculated isomerization intermediates, and (3) either H or X migration to form haloethyl-
barrier HBrCBr  Br.61 Thus, the pathway to the Br2 product idene, followed by three-center elimination. Among these
mainly proceeds via this lower isomerization TS. mechanisms, molecular halogen elimination has never been
A similar isomerization/photodissociation process has been reported except for 1,2-dibromoethylene but yielding the skep-
found in dihalomethanes in the condensed phase. Phillips tical results with the PTS technique.21 In contrast, as with the
and co-workers demonstrated that iso-CH2I–Br is mainly cases of bromoalkanes, the CRDS spectra of Br2 (v = 0 and 1)
responsible for the transient absorption spectrum following can be readily acquired at 248 nm in the isomeric forms of
either A-band or B-band excitation of bromoiodomethane in 1,1- and 1,2-dibromoethylenes.70,71 The spectrum obtained
cyclohexane solution in the nanosecond resonance Raman in 1,1-C2H2Br2 is given as an example (Fig. 7). All possible
experiments.63 The UV excitation leads to a fast C–I or C–Br examinations were carried out to ensure that the obtained Br2
bond fission, followed by recombination to form iso-CH2Br–I or fragments stem from the primary photodissociation occurring
iso-CH2I–Br within the solvent cage. But only the latter photo- on the ground state surface.
product can be detected during the nanosecond period, As dibromoethylenes are excited to the (p, p*) state at
because iso-CH2I–Br species is stable with a longer lifetime 248 nm, the mutually twisted BrCBr (or HCBr) and HCH (or
than iso-CH2Br–I. In contrast, the former photoproduct was mainly HCBr) groups may enhance the electronic coupling to facilitate
observed under solid matrix at 12 K following the A-band the IC process.72 The ab initio potential energy calculations help
excitation,64,65 since the elongated C–I bond breaks rapidly understand the detailed photodissociation mechanisms. As
into CH2Br + I, followed by recombination due to the cage effect. shown in Fig. 8, 1,1-dibromoethylene follows essentially four

This journal is © the Owner Societies 2014 Phys. Chem. Chem. Phys., 2014, 16, 7184--7198 | 7189
View Article Online

PCCP Perspective
Published on 21 February 2014. Downloaded by Academia Sinica - Taipei on 15/12/2017 09:16:31.

Fig. 5 The dissociation channels of 1,2-C2H4Br2, in which the energies in kJ mol1 relative to 1,2-C2H4Br2 are computed with CCSD(T)/cc-pVTZ level of
theory with B3LYP/6-311G(d,p) zero-point energy corrections at B3LYP/6-311G(d,p) optimized geometries. The vertical energies of 1,2-C2H4Br2 excited
states are computed with MRDCI/cc-pVTZ at B3LYP/6-311G(d,p) optimized ground state geometry.

Fig. 6 The dissociation channels of 1,1-C2H4Br2, in which the energies in kJ mol1 relative to 1,1-C2H4Br2 are computed with the methods identical to
the calculations for 1,2-C2H4Br2 in Fig. 5.

pathways to the products of HCCH + Br2.71 The route (a) proceeds in which TSi3-Br2(b-2a) is an isomerization TS lying at
along i4 (1,1-C2H2Br2) - TSi4-Br2 - HCCH + Br2. The route (b) 105 kJ mol1 less than TScis-Br2(b-2b). The route (c) is along
follows i4 - TSi2-i4 - TSi3-Br2(b-2a) or TScis-Br2(b-2b) - HCCH + Br2, i4 - TSi3-i4 TSi3-Br2 - HCCH + Br2, sharing the same isomerization

7190 | Phys. Chem. Chem. Phys., 2014, 16, 7184--7198 This journal is © the Owner Societies 2014
View Article Online

Perspective PCCP
Published on 21 February 2014. Downloaded by Academia Sinica - Taipei on 15/12/2017 09:16:31.

Fig. 7 The CRDS spectra of Br2 v = 0 and 1 obtained from photolysis of 1,1-C2H2Br2 at a pressure of 0.7 Torr. Spectral simulation takes into account
the population (a) only in the v = 0 level, (b) only in the v = 1 level, and (c) in both the v = 1 and 0 levels with a branching ratio of 0.55, in comparison
with (d) the experimental results.

Fig. 8 The Br2 dissociation channels of 1,1-C2H2Br2 on its adiabatic singlet ground state PES, in which the energies in kJ mol1 relative to 1,1-C2H2Br2 are
computed with CCSD(T)/cc-pVTZ level of theory with B3LYP/6-311G(d,p) zero-point energy corrections.

TS with the route (b-2a). Accordingly, the three-center elimination energy conservation throughout the reaction, for an unimolecular
of Br2 via TSi3-Br2 along the routes (b-2a) and (c) are energetically k
reaction A ! Aa ! P, where A* is the energized reactant, Aa
more favored. The Br–Br bond distance in the TSi3-Br2 structure is the transition state, and P the products, the rate constant k(E)
largely stretched, thereby Br2 may be produced vibrationally as a function of energy is expressed as
hot. Similarly, in 1,2-C2H2Br2 the three-center mechanism via
isomerization TSi3-Br2 to produce vibrationally hot Br2 prevails s W a ðE  E a Þ
kðEÞ ¼ ; (5)
over a four-center elimination from the cis-isomer.70 The h rðEÞ
former pathway shows a dissociation rate constant about
30 times faster than the latter calculated by the Rice–Ramsperger– where s is the symmetry factor, Wa the number of states of the
Kassel–Marcus (RRKM) method.70,73 TS, and r the density of states of the reactant. The saddle-point
The RRKM rate constants were calculated on singlet ground method74,75 is applied to evaluate the number of states and
state (or the first triplet state) PESs of the molecules studied. If the density of states; the molecule is viewed as a collection
the rate of the energy equilibration is faster than the reaction of harmonic oscillators. The geometries and the harmonic
rate, the rate constant can be explained statistically. Given total frequencies of the reactant, TS, and products were essentially

This journal is © the Owner Societies 2014 Phys. Chem. Chem. Phys., 2014, 16, 7184--7198 | 7191
View Article Online

PCCP Perspective

obtained at the level of the hybrid density functional theory, appears to have a structure similar to the isomerization TS that
B3LYP/6-311G(d,p); the energies are further refined by the carries a low energy barrier to form the halogen molecules, as
coupled cluster CCSD(T)/cc-pVTZ with B3LYP/6-311G(d,p) described above.61–65 For the Br2 product generated from this
zero-point energy corrections.71 process, the related laser energy and pressure dependence
For some cases, the Br2 production may not be sufficiently each is expected to yield a slope of unity that agrees with our
interpreted by the pathways constructed and the subsequent CRDS findings. Therefore, partial contribution of Br2 from
RRKM calculations based on the statistical evaluation. this roaming pathway might account for a larger quantum
CH2BrCOBr as an example gave rise to the Br2 yield of 0.24  yield observed.
0.08 in photodissociation at 248 nm,27 about 10 times larger
Published on 21 February 2014. Downloaded by Academia Sinica - Taipei on 15/12/2017 09:16:31.

than that by the RRKM calculations while taking into account 3.4. Photodissociation of I-substituted methane
the other competitive dissociation channels. Therefore, an Diiodomethane (CH2I2) as an example in photodissociation of
alternative mechanism might be proposed such as roaming I-substituted methane contains at least three dissociation
pathway.76–84 In this photodissociation process, recoiling atoms channels of CHI + I(2P3/2), CHI + I*(2P1/2) and CH2 + I2 in the
or radicals, weakly bound to the other moiety, can meander and UV excitation. The lowest UV absorption spectra of CH2I2 split
finally undergo intramolecular abstraction forming molecular into five overlapping states with the 1B1, 2B1, 1B2, 1A1, and 2A1
products. Based on the dissociation pathways (Fig. 9),27 a symmetry, in the order of increasing energy.85–88 Among them,
fraction of CH2BrCOBr molecules may undergo a radical dis- the 1B1 and 2A1 states are correlated with the dissociation
sociation producing CH2BrCO + Br at 292.5 kJ mol1 via channel of CH2I + I(2P3/2), while the remaining states are all
internal conversion. The barrierless radical channel lies in an correlated with CH2I + I*(2P1/2).88 CH2I2 is dominated by the
energy state in proximity to the TS at 294.0 kJ mol1 associated dissociation channel of CH2I and I (or I*). Thus, its photo-
with the route to Br2 + CH2CO; such a condition is essentially dissociation dynamics is actively focused on determination of
favored for the roaming occurrence. The Br atom, carrying quantum yields of released iodine atoms,86,89 translational
a small translational energy released from this barrierless energy distribution,85,87,88 internal energy deposition in the
channel, may then be feasibly attracted around a wide flat radical,86 and anisotropy parameter measurements.85,87,88
region of the Br  BrCH2CO PES prior to abstraction of another Thus far, the I2 elimination from CH2I2 was reported in the
Br atom on CH2BrCO. Such a roaming saddle point Br  BrCH2CO vacuum ultraviolet region via either single- or multi-photon

Fig. 9 The dissociation pathways of CH2BrCOBr, in which the energies in units of kJ mol1 relative to the ground state are computed with CCSD(T)/
cc-pVTZ level of theory with B3LYP/cc-pVTZ zero-point energy corrections at B3LYP/cc-pVTZ optimized geometries.

7192 | Phys. Chem. Chem. Phys., 2014, 16, 7184--7198 This journal is © the Owner Societies 2014
View Article Online

Perspective PCCP

excitation, and only the excited states of I2 were observed in the regression coefficient, R2 = 0.98. The I2 product was further
emission measurements. By using a Kr(1165, 1236 Å) lamp, verified to result from a primary dissociation channel following
Okabe et al.90 obtained a primary fragment I2(3P2g) that was one-photon absorption of CH2I2 at 248 nm.93 The quantum
vibrationally excited up to v = 35 with a quantum yield of yield f of the I2 elimination channel can be determined
0.006 at 1236 Å, but little vibrational excitation was found at by comparison with a reference sample of CH2Br2 by the
1306 Å. Donovan and coworkers91 observed fluorescence from following equation:93
F1Su+ and D1Su+ states by using multiphoton excitation at fI2 ½I2 sCH2 Br2 nCH2 Br2
248 and 193 nm. Marvet and Dantus92 verified a concerted ¼ : (6)
fBr2 ½Br2 sCH2 I2 nCH2 I2
elimination of I2(3P2g) in less than 100 fs by two-photon
Published on 21 February 2014. Downloaded by Academia Sinica - Taipei on 15/12/2017 09:16:31.

excitation at 310 nm using pump–probe femtosecond spectro-


scopy. The ground state I2 product with the ultraviolet excita- Given the halogen concentrations determined by eqn (3), the
tion has never been observed, although this channel was quantum yield of Br2,56 the absorption cross sections of sCH2I2
suspected to probably exist by Leone and co-workers.86 and sCH2Br2 equal to 1.6  1018 (ref. 96) and 3.7  1019 cm2
The ground state I2 product and dissociation mechanism in (ref. 56) at 248 nm, respectively, and the intensity ratio of the
a single-photon dissociation of CH2I2 at 248 nm was character- halogen molecules along with the individual absorption cross
ized for the first time by using CRDS.93 Since the CRDS section, the quantum yield of I2 dissociated from CH2I2 is
experiments were carried out at room temperature, a cluster evaluated to be 0.0040  0.0025. The obtained quantum yield
formation like (CH2I2)2 is negligible. Such a cluster-induced happens to be as small as those of the excited state I2,90 but
photochemistry was investigated in CH3I prepared in a jet- they have different dissociation pathways.
cooled temperature, thereby leading to the I2 product following Fig. 11 shows two pathways for the I2 elimination. Route (1)
(CH3I)2 photodissociation.94 CH3I was also found to produce encounters a TS (tsI2) along the PES at 407.0 kJ mol1 above the
electronically excited B states of I2 on a MgO(100) surface, ground state CH2I2, while route (2) goes through a TS (tsCI) at
due to multi-photon excitation followed by the reaction of 191.5 kJ mol1 and an intermediate state at 179.3 kJ mol1
I* + I - I2(B) on the surface.95 As shown in Fig. 10, a portion before reaching the products of I2 + CH2. The tsI2 structure
of CRDS spectra of I2 (v = 0,1,2) in the B3Pou+ ’ X1Sg+ shows the character of a C1 symmetry, with the I–C–I angle
transition in photolysis of CH2I2 at 248 nm contains the v = suppressed to 60.21 and the C–I bond lengths elongated asym-
0, 1, and 2 bands corresponding to (39,0)–(56,0), (44,1)–(62,1), metrically to 2.578 and 3.017 Å, such that the associated I–I
and (49,2)–(50,2).32 Because the I2 product is heavy and easily moiety may be released. Another tsCI structure has a Cs
stick onto the chamber wall, such experiments were conducted symmetry with the I–C–I angle of 70.81 and the two C–I bond
with an increased pumping speed to make sure no memory lengths changed to 1.991 and 3.460 Å. The tsCI structure is
effect from the residual I2. The CRDS spectrum is accompanied similar to an isomerization TS in which a X–X bond is formed
by a simulated counterpart based on a similar equation as before reaching the products, like the cases of CF2Cl2 and
in eqn (1) (Fig. 10). The spectral positions and rotational CF2Br2.57,61 Note that in this work the geometries and the
intensities are comparable between them, although the signal harmonic frequencies of the reactant, intermediate, TS, and
quality is worsened by the background noise. Accordingly, products are obtained at the level of B3LYP/midix,97 instead
the vibrational population ratio of v(0) : v(1) : v(2) is determined of B3LYP/6-311G(d,p) adopted in the Br-related systems; the
to be 1 : (0.65  0.10) : (0.30  0.05), corresponding to a Boltzmann- energies are further refined by the coupled cluster CCSD(T)/
like vibrational temperature of 544  73 K with a square of midix with B3LYP/midix zero-point energy corrections. The
attempts to locate the TS for I elimination by B3LYP were not
successful, the multi-configuration approach, CASSCF/midix,
was employed instead. The basis set midix can be appropriately
utilized for a moderate size like I, and succeed to predict a
reasonable I2 bond length. According to the RRKM method, the
dissociation rate constants for the routes (1) and (2) are
calculated to be 2.41  109 and 3.52  1011 s1, respectively.
Thus, I2 is favored to eliminate via an asynchronous three-
center mechanism. As compared to the I2 equilibrium bond
length of 2.727 Å, the dominant pathway via tsCI with a I  I
distance of 3.379 Å is anticipated to produce vibrationally
excited I2, in agreement with the observation.
As compared to CH2Br2, the excited states of CH2I2 are
stabilized to cause a red shift in the excitation transition. Thus,
CH2I2 is excited at 248 nm to the mixed state of 1B2 and 1A1,
Fig. 10 A portion of I2 spectra (v = 0, 1, and 2) acquired in the photolysis
whereas CH2Br2 reaches a lower mixed state of B1 and B2.56,98,99
of CH2I2 at 248 nm. (a) denotes trace acquired experimentally, while (b) is a The 1A1 and 1B2 symmetries become A 0 and A00 , respectively, as
simulated spectrum. the C–I bond is elongated and the CH2I2 structure changes to

This journal is © the Owner Societies 2014 Phys. Chem. Chem. Phys., 2014, 16, 7184--7198 | 7193
View Article Online

PCCP Perspective
Published on 21 February 2014. Downloaded by Academia Sinica - Taipei on 15/12/2017 09:16:31.

Fig. 11 The dissociation channels initiated from the ground state potential energy surface of CH2I2, in which the energies in kJ mol1 relative to CH2I2
are computed with CCSD(T)/midix level of theory with B3LYP/midix zero-point energy corrections at B3LYP/midix optimized geometries.

the Cs symmetry. In the IC process, the 1A 0 –X̃1A 0 coupling is vibrational states for CH2I2 and CH2Br2 at 248 nm were thus
strengthened to facilitate the level-to-level interaction, while the calculated to yield 1.15  106 and 5.61  105 cm1, respectively,
interaction between the 1A00 and X̃1A 0 states with different in which the harmonic frequencies were calculated by B3LYP/
symmetry is relatively weak. Accordingly, the population lying midix.93 As temperature increases, the density of states of CH2I2
in the 1A 0 state is favored to transit to the high vibrational levels increases more rapidly to make the IC process more efficient,
of the ground state X̃1A 0 prior to direct photodissociation. The with respect to CH2Br2. That might cause a large discrepancy in
photodissociation of CH2I2 and CH2Br2 behaves differently the temperature effect.
in some aspects. For instance, the quantum yield of the I2
elimination in CH2I2 is 0.0040  0.0025, in contrast to a result 3.5. Photodissociation of thionyl chloride and sulfuryl
of 0.2  0.1 for the Br2 elimination in CH2Br2. The reasons are chloride
yet to be known. However, the fraction of population under- Thionyl chloride (SOCl2) and sulfuryl chloride (SO2Cl2) have
going internal conversion causes the difference between them. become of atmospheric interest as they may eliminate both Cl
For CH2I2, only the 1A1 population is allowed to undergo IC, and SO; the latter pollutant is readily oxidized to H2SO4, the
sharing a small fraction in the excitation band. As for CH2Br2, acid rain source. The photochemistry of SOCl2 has been well
the population excited to the B1 state takes a larger fraction. investigated, producing three dissociation channels: (1) SOCl +
Like the A1 state, B1 is efficient to undergo IC, as the Br–CH2Br Cl, (2) SO + Cl2, and (3) SO + 2Cl. The insight into photolysis-
bond is elongated. The temperature effect is another different wavelength dependence of dissociation channels with the
behavior between these two structural analogs. The I2 intensity related branch ratios,12,100–102 anisotropy parameter for each
increases up to 220%, but the Br2 intensity increases only about channel,12,13,103 dissociation characterization for the three-
16% within a similar temperature increment of 24–30 K. The body production,12,104 and nascent vibrational and spin-state
heavier iodine mass should result in a larger density of states distribution of SO103,105 was mainly gained. The prior studies
for CH2I2 at the same excitation energy. The density of are essentially focused on probing of the Cl (Cl*) and SO

7194 | Phys. Chem. Chem. Phys., 2014, 16, 7184--7198 This journal is © the Owner Societies 2014
View Article Online

Perspective PCCP

no contribution of such a three-body dissociation was observed


when the wavelength was changed to 248 nm.103 The atomic Cl
recombination from this dissociation channel is thus
neglected. Further, the reaction of eliminated Cl with SOCl2
requires two SOCl2 molecules that are opposed to the pressure
dependence measurement. Because the Cl yield was reported to
be 96.5% at 248 nm,103 a pseudo-first order condition cannot
be applied in this case. These experimental results may exclude
the probable contributions from atomic Cl recombination,
Published on 21 February 2014. Downloaded by Academia Sinica - Taipei on 15/12/2017 09:16:31.

secondary reaction of energized fragments, or reactions


between Cl and the precursor. Determination of quantum yield
of the Cl2 channel is underway in our group.
To the best of our knowledge, photochemistry of SO2Cl2 has
never been investigated except for photolysis-wavelength
dependence of absorption cross sections105 which turns out
Fig. 12 A portion of Cl2 spectra (v = 0) acquired in the photolysis of SOCl2
and SO2Cl2 at 248 nm. A 5% of Cl2 gas in Ne buffer gas and simulated to be larger by about one order of magnitude than SOCl2.
counterpart are also included for comparison. According to the estimate with heat of formation data and
atomic energy levels,105 the following dissociation channels are
anticipated depending on the photolysis wavelength: (1) O2S +
fragments for understanding the dynamical complexity especially Cl2, (2) O2SCl + Cl (Cl*), (3) O2S + 2Cl (Cl*), (4) OSCl + OCl, (5)
for the channels (1) and (3). Nevertheless, whether the Cl2 fragment OSCl2 + O(3P) (O(1D)), and (6) OSCl + O(3P) + Cl. None of them
is produced is suspected by a recent investigation using three- have been investigated. As shown in Fig. 12, this account
dimensional imaging to probe photofragments at 235 nm.13 reports the first case of the Cl2 optical spectra of the channel (1)
Thus far, the Cl2 fragment was monitored using time-of- in photolysis at 248 nm using the CRDS technique. The required
flight mass spectrometer,12,13 but its optical spectrum has not experiments were also carried out to confirm that the Cl2 product
been observed for understanding the internal state distribu- is obtained via primary fragmentation. Further characterization of
tions. In this account, a photolysis laser beam at 248 nm is the photodissociation dynamics is still underway.
applied to excite SOCl2 in the (pSCl*, nS) transition with an
absorption cross section of 7.1  1018 cm1.106 Fig. 12 shows
the optical spectrum of Cl2 (v = 0) following photolysis of SOCl2 4. Conclusion
at a pressure of 175 mTorr that is detected for the first time
using the CRDS method. The spectra of Cl2 (v = 0,1) and Molecular halogen elimination has been seldom reported
(v = 0,1,2) were also obtained in the region of 503–504 and among the photodissociation channels of halogen-containing
511–513 nm, respectively. When the photolysis laser is turned compounds ever studied. This account is intended to attract the
off, the Cl2 signal disappears. A pure Cl2 gas diluted to 5% in attention of this molecular channel by reviewing the CRDS
Ne buffer gas at a total pressure of 165 mTorr is compared detection of the X2 (Br2, Cl2 and I2) fragments resulting from a
in Fig. 12. The spectral consistency indicates that the Cl2 variety of multi-X-containing molecules with characterization
spectrum dissociated is not interfered with by other dissocia- of the related elimination pathways. Such a channel has been
tion products. According to eqn (1), a simulated counterpart is verified to stem from primary dissociation of a single molecule
calculated for comparison, in which the isotopic variants on the ground state PES via internal conversion. Like the
37 37
Cl Cl and 35Cl37Cl are not taken into account. From the halogen products studied, some molecular channels are found
simulation, the vibrational population ratio at each state may to dissociate similarly from the ground-state PESs following
be analyzed. photodissociation.107,108 According to the ab initio potential
To remove the possibility of Cl2 resulting from atomic Cl energy calculations, the parent molecules in the highly
recombination, a single Cl-containing molecule is substituted vibrational levels of the ground state are energetically accessi-
such as acetyl chloride (CH3COCl) with a comparable absorp- ble to surpass the transition states and break into the X2
tion cross section, but no any Cl2 signal can be detectable products. Even for the simple halomethanes, their photodisso-
even if the sample pressure is increased to 600 mTorr. The ciation behaves distinctly from three-center to four-center
pressure and laser energy dependence of the Cl2 fragment are concerted and from asynchronous to sequential mechanisms.
also measured, yielding a straight line with a slope of unity, As compared to bromomethanes and bromoethanes, the photo-
indicative of a single-photon involvement in a single molecular dissociation mechanisms of dibromoethylenes are more com-
elimination. Two photons are otherwise required for the plicated; more than one route may be followed to release the
secondary recombination process, if the source of Cl atoms is Br2 fragment. With respect to X elimination, the obtained X2
from two separate SOCl2 molecules. A concerted three-body yield is low ranging from 0.004 to 0.3. However, the small
dissociation (SO + Cl + Cl) was reported to be the main channel amount of halogen molecules might have a chance to become
accounting for 480% of the overall process at 193 nm, whereas or induce a stable reservoir in the stratosphere. The ozone loss

This journal is © the Owner Societies 2014 Phys. Chem. Chem. Phys., 2014, 16, 7184--7198 | 7195
View Article Online

PCCP Perspective

rate is often used as an indicator of the extent of stratospheric 19 M. Chandra, D. Senapati, M. Tak and P. K. Das, Chem.
ozone depletion along specific altitudes and latitudes. This rate Phys. Lett., 2006, 430, 32–35.
is defined as the ratio of reactive free radicals (such as Cl, ClO, 20 J. F. Riehl, D. G. Musaev and K. Morokuma, J. Chem. Phys.,
Br, and BrO) to the gas sources and halogen-induced reservoirs 1994, 101, 5942–5956.
(such as HX and XONO2) that reach the stratosphere.5,7 Under- 21 Y. R. Lee, C. C. Chou, Y. J. Lee, L. D. Wang and S. M. Lin,
standing the X2 dissociation mechanism should contribute to J. Chem. Phys., 2001, 115, 3195–3200.
assessing the issues of halogen-induced ozone destruction and 22 L. Hua, W. B. Lee, M. H. Chao, B. Zhang and K. C. Lin,
halogen-related environmental change. J. Chem. Phys., 2011, 134, 194312.
23 W. Shi, V. N. Staroverov and R. H. Lipson, J. Chem. Phys.,
Published on 21 February 2014. Downloaded by Academia Sinica - Taipei on 15/12/2017 09:16:31.

2009, 131, 154304.


Acknowledgements 24 L. Hua, W. B. Lee, M. H. Chao, B. Zhang and K. C. Lin,
J. Chem. Phys., 2011, 134, 194312.
This work is supported by National Taiwan University, Ministry
25 Y. J. Jee, Y. J. Jung and K. H. Hung, J. Chem. Phys., 2001,
of Education, and National Science Council of Taiwan, Republic
115, 9739–9746.
of China, under contract no. NSC 99-2113-M-001-025-MY3.
26 J. Tellinghuisen, J. Chem. Phys., 2001, 115, 10417–14024.
27 H. Fan, P. Y. Tsai, K. C. Lin, C. W. Lin, C. Y. Yan, S. W. Yang
and A. H. H. Chang, J. Chem. Phys., 2012, 137, 214304.
References
28 J. J. Scherer, J. B. Paul, A. O’Keefe and R. J. Saykally, Chem.
1 M. J. Molina and F. S. Rowland, Nature, 1974, 249, 810–812. Rev., 1997, 97, 25–51.
2 J. Farman, B. Gardiner and J. Shanklin, Nature, 1985, 315, 29 P. Y. Wei, Y. P. Chang, Y. S. Lee, W. B. Lee, K. C. Lin,
207–210. K. T. Chen and A. H. H. Chang, J. Chem. Phys., 2007,
3 S. A. Montzka, J. H. Butler, R. C. Myers, T. M. Thompson, 126, 034311.
T. H. Swanson, A. D. Clarke, L. T. Lock and J. W. Elkins, 30 H. L. Lee, P. C. Lee, P. Y. Tsai, K. C. Lin, H. H. Kuo,
Science, 1996, 272, 1318–1322. P. H. Chen and A. H. H. Chang, J. Chem. Phys., 2009,
4 S. Montzka, J. Butler, R. Myers, T. Thompson, T. Swanson, 130, 184308.
A. Clarke, L. Lock and J. Elkins, Science, 1996, 272, 31 C. C. Wu, H. C. Lin, Y. B. Chang, P. Y. Tsai, Y. Y. Yeh,
1318–1322. H. Fan, K. C. Lin and J. S. Francisco, J. Chem. Phys., 2012,
5 S. Solomon, R. R. Garcia and A. R. Ravishankara, 135, 234308.
J. Geophys. Res., 1994, 99, 20491–20499. 32 Laboratoire Aime Cotton, http://www.lac.u-psud.fr/-Atlas-
6 M. Bilde, T. J. Wallington, C. Ferronato, J. J. Orlando, de-l-iode-et-du-tellure-.
G. S. Tyndall, E. Estupinan and S. Haberkorn, J. Phys. 33 J. A. Coxon and R. Shanker, J. Mol. Spectrosc., 1978, 69,
Chem. A, 1998, 102, 1976–1986. 109–122.
7 J. S. Daniel, S. Solomon, R. W. Portmann and R. R. Garcia, 34 A. Mellouki, R. K. Talukdar, A.-M. Schmoltner, T. Gierczak,
J. Geophys. Res., 1999, 104, 23871–23880. M. J Mills, S. Solomon and A. R. Ravishankara, Geophys.
8 D. E. Mann and B. A. Thrush, J. Chem. Phys., 1960, 33, Res. Lett., 1992, 19, 2059–2062.
1732–1734. 35 Y. R. Lee, C. C. Chen and S. M. Lin, J. Chem. Phys., 2003,
9 C. L. Sam and J. T. Yardley, Chem. Phys. Lett., 1979, 61, 118, 10494–10501.
509–512. 36 U. Marvet, E. J. Brown and M. Dantus, Phys. Chem. Chem.
10 D. Krajnovich, Z. Zhang, L. Butler and Y. T. Lee, J. Phys. Phys., 2000, 2, 885–891.
Chem., 1984, 88, 4561–4566. 37 U. Marvet, Q. Zhang and M. Dantus, J. Phys. Chem. A, 1998,
11 M. R. Cameron, S. A. Johns, G. F. Metha and S. H. Kable, 102, 4111–4117.
Phys. Chem. Chem. Phys., 2000, 2, 2539–2547. 38 P. Zou, J. Shu, T. J. Sears, G. E. Hall and S. W. North, J. Phys.
12 G. Baum, C. S. Effenhauser, P. Felder and J. R. Huber, Chem. A, 2004, 108, 1482–1488.
J. Phys. Chem., 1992, 96, 756–764. 39 D. Xu, J. S. Francisco, J. Huang and W. M. Jackson, J. Chem.
13 A. Chichinin, T. S. Einfeld, K.-H. Gericke and J. Grunenberg, Phys., 2002, 117, 2578–2585.
Phys. Chem. Chem. Phys., 2005, 7, 301–309. 40 F. Taketani, K. Takahashi and Y. Matsumi, J. Phys. Chem. A,
14 K. Sato, S. Tsunashima, T. Takayanagi, G. Fujisawa and 2005, 109, 2855–2860.
A. Yokoyama, J. Chem. Phys., 1997, 106, 10123–10133. 41 L. J. Butler, E. J. Hintsa and Y. T. Lee, J. Chem. Phys., 1986,
15 D. A. Blank, W. Sun, A. G. Suits, Y. T. Lee, S. W. North and 84, 4104–4106.
G. E. Hall, J. Chem. Phys., 1998, 108, 5414–5425. 42 L. J. Butler, E. J. Hintsa and Y. T. Lee, J. Chem. Phys., 1987,
16 C. Y. Wu, C. Y. Chung, Y. C. Lee and Y. P. Lee, J. Chem. 86, 2051–2074.
Phys., 2002, 117, 9785–9792. 43 R. F. Barrow, T. C. Clark, J. A. Coxon and K. K. Yee,
17 S. H. Lee, W. K. Chen, C. Chaudhuri, W. J. Huang and J. Mol. Spectrosc., 1974, 51, 428–449.
Y. T. Lee, J. Chem. Phys., 2006, 125, 144315. 44 S. Gerstenkorn and P. Luc, J. Phys., 1989, 50, 1417–1432.
18 G. X. He, Y. A. Yang, Y. Huang and R. J. Gordon, J. Phys. 45 J. A. Coxon, J. Quant. Spectrosc. Radiat. Transfer, 1972, 12,
Chem., 1993, 97, 2186–2193. 639–650.

7196 | Phys. Chem. Chem. Phys., 2014, 16, 7184--7198 This journal is © the Owner Societies 2014
View Article Online

Perspective PCCP

46 G. Herzberg, Molecular Spectra and Molecular Structure: I. 72 I. Ohmine, J. Chem. Phys., 1985, 83, 2348–2362.
Spectra of Diatomic Molecules, van Nostrand Reinhold, 73 A. H. H. Chang, D. W. Hwang, X. M. Yang, A. M. Mebel,
Toronto, 1950, pp. 133–139. S. H. Lin and Y. T. Lee, J. Chem. Phys., 1999, 110,
47 R. C. Sharma, H. Y. Huang, W. T. Chuang and K. C. Lin, 10810–10820.
Spectrochim. Acta, Part A, 2005, 60, 2115–2120. 74 A. H. H. Chang, A. M. Mebel, X. M. Yang, S. H. Lin and
48 H. Y. Huang, W. T. Chuang, R. C. Sharma, C. Y. Hsu, Y. T. Lee, J. Chem. Phys., 1998, 109, 2748–2761.
K. C. Lin and C. H. Hu, J. Chem. Phys., 2004, 121, 5253–5260. 75 H. Eyring, S. H. Lin and S. M. Lin, Basic Chemical Kinetics,
49 B. Abel, H. Hippler, N. Lange, J. Schuppe and J. Troe, Wiley, New York, 1980.
J. Chem. Phys., 1994, 101, 9681–9690. 76 D. Townsend, S. A. Lahankar, S. K. Lee, S. D. Chambreau,
Published on 21 February 2014. Downloaded by Academia Sinica - Taipei on 15/12/2017 09:16:31.

50 V. L. Orkin and E. E. Kasimovskaya, J. Atmos. Chem., 1995, A. G. Suits, X. Zhang, J. Rheinecker, L. B. Harding and
21, 1–11. J. M. Bowman, Science, 2004, 306, 1158–1161.
51 S. L. Baughcum and S. R. Leone, J. Chem. Phys., 1980, 72, 77 A. G. Suits, Acc. Chem. Res., 2008, 41, 873–881.
6531–6545. 78 H. M. Yin, S. H. Kable, X. Zhang and J. M. Bowman,
52 A. T. J. B. Eppink and D. H. Parker, J. Chem. Phys., 1998, Science, 2006, 311, 1443–1446.
109, 4758–4767. 79 P. L. Houston and S. H. Kable, Proc. Natl. Acad. Sci. U. S. A.,
53 L. J. Butler, Annu. Rev. Phys. Chem., 1998, 49, 125–171. 2006, 103, 16079–16082.
54 S. Enami, S. Hashimoto, M. Kawasaki, Y. Nakano, 80 B. R. Heazlewood, M. J. T. Jordan, S. H. Kable, T. M. Selby,
T. Ishiwata, K. Tonokura and T. J. Wallington, J. Phys. D. L. Osborn, B. C. Shepler, B. J. Braams and
Chem. A, 2005, 109, 1587–1593. J. M. Bowman, Proc. Natl. Acad. Sci. U. S. A., 2008, 105,
55 J. Sehested, M. Bilde, T. Mogelberg, T. J. Wallington and 12719–12724.
O. J. Nielsen, J. Phys. Chem., 1996, 100, 10989–10998. 81 S. Maeda, K. Ohno and K. Morokuma, J. Phys. Chem. Lett.,
56 P. Y. Wei, Y. P. Chang, W. B. Lee, Z. Hu, H. Y. Huang, 2010, 1, 1841–1845.
K. C. Lin, K. T. Chen and A. H. H. Chang, J. Chem. Phys., 82 M. P. Grubb, M. L. Warter, A. G. Suits and S. W. North,
2006, 125, 133319. J. Phys. Chem. Lett., 2010, 1, 2455–2458.
57 C. Y. Hsu, H. Y. Huang and K. C. Lin, J. Chem. Phys., 2005, 83 C. Chen, B. Braams, D. Y. Lee, J. M. Bowman, P. L.
123, 134312. Houston and D. Stranges, J. Phys. Chem. Lett., 2010, 1,
58 F. Jensen, Introduction to computational chemistry, Wiley, 1875–1880.
New York, 2nd edn, 2007, p. 136. 84 M. H. Chao, P. Y. Tsai and K. C. Lin, Phys. Chem. Chem.
59 J. Zheng, J. R. Gour, J. J. Lutz, M. Wloch, P. Piecuch and Phys., 2011, 13, 7154–7161.
D. G. Truhlar, J. Chem. Phys., 2008, 128, 044108. 85 M. Kawasaki, S. J. Lee and R. Bersohn, J. Chem. Phys., 1975,
60 T. Noro, M. Sekiya and T. Koga, Theor. Chem. Acc., 2012, 63, 809–814.
131, 1124–1128 (basis set is available from http://setani.sci. 86 S. L. Baughcum and S. R. Leone, J. Chem. Phys., 1980, 72,
hokudai.ac.jp/sapporo/Welcome.do). 6531–6545.
61 A. Kalume, L. George and S. A. Reid, J. Phys. Chem. Lett., 87 K. W. Jung, T. S. Ahmadi and M. A. El-Sayed, Bull. Korean
2010, 1, 3090–3095. Chem. Soc., 1997, 18, 1274–1280.
62 D. Krajnovich, F. Huisken, Z. Zhang, Y. R. Shen and 88 H. Xu, Y. Guo, S. Liu, X. Ma, D. Dai and G. Sha, J. Chem.
Y. T. Lee, J. Chem. Phys., 1982, 77, 5977–5989. Phys., 2002, 117, 5722–5729.
63 X. Zheng and D. L. Phillips, J. Chem. Phys., 2000, 113, 89 J. B. Koffend and S. R. Leone, Chem. Phys. Lett., 1981, 81,
3194–3203. 136–141.
64 G. Maier and H. P. Reisenauer, Angew. Chem., Int. Ed. Engl., 90 H. Okabe, M. Kawasaki and Y. Tanaka, J. Chem. Phys.,
1986, 25, 819–822. 1980, 73, 6162–6166.
65 G. Maier, H. P. Reisenauer, J. Lu, L. J. Scaad and B. A. Hess, 91 C. Fotakis, M. Martin and R. J. Donovan, J. Chem. Soc.,
Jr., J. Am. Chem. Soc., 1990, 112, 5117–5122. Faraday Trans. 2, 1982, 78, 1363.
66 J. J. Lin, S. M. Wu, D. W. Hwang, Y. T. Lee and X. Yang, 92 U. Marvet and M. Dantus, Chem. Phys. Lett., 1996, 256,
J. Chem. Phys., 1998, 109, 10838–10846. 57–62.
67 Y. B. Huang and R. J. Gordon, J. Chem. Phys., 1997, 106, 93 S. Y. Chen, P. Y. Tsai, H. C. Lin, C. C. Wu, K. C. Lin,
1418–1420. B. J. Sun and A. H. H. Chang, J. Chem. Phys., 2011,
68 J. Y. Tu, J. J. Lin, Y. T. Lee and X. M. Yang, J. Chem. Phys., 134, 034315.
2002, 116, 6982–6989. 94 Y. B. Fan and D. J. Donaldson, J. Chem. Phys., 1992, 97,
69 G. E. Hall, J. T. Muckerman, J. M. Preses, R. E. Weston, 189–196.
G. W. Flynn and A. Persky, J. Chem. Phys., 1994, 101, 95 M. E. Vaida and T. M. Bernhardt, Faraday Discuss., 2012,
3679–3688. 157, 437–449.
70 Y.-P. Chang, P.-C. Lee, K.-C. Lin, C.-H. Huang, B. J. Sun and 96 J. C. Mossinger, D. E. Shallcross and R. A. Cox, J. Chem.
A. H.-H. Chang, ChemPhysChem, 2008, 9, 1137–1145. Soc., Faraday Trans., 1998, 94, 1391–1396.
71 P.-C. Lee, P.-Y. Tsai, M.-K. Hsiao, K.-C. Lin, C.-H. Huang 97 R. E. Easton, D. J. Giesen, A. Welch, C. J. Cramer and
and A. H.-H. Chang, ChemPhysChem, 2009, 10, 672–679. D. G. Truhlar, Theor. Chim. Acta, 1996, 93, 281–301.

This journal is © the Owner Societies 2014 Phys. Chem. Chem. Phys., 2014, 16, 7184--7198 | 7197
View Article Online

PCCP Perspective

98 G. C. Causley and B. R. Russell, J. Chem. Phys., 1975, 62, 104 H. Wang, X. Chen and B. R. Weiner, J. Phys. Chem., 1993,
848–857. 97, 12260–12268.
99 H. Y. Xiao, Y. J. Liu, J. G. Yu and W. H. Fang, J. Comput. 105 X. Chen, F. Asmar, H. Wang and B. R. Weiner, J. Phys.
Chem., 2008, 29, 2513–2519. Chem., 1991, 95, 6415–6417.
100 H. Okabe, J. Am. Chem. Soc., 1971, 93, 7095–7096. 106 A. P. Uthman, P. J. Demlein, T. D. Allston, M. C. Withiam,
101 H. Kanamori, E. Tiemann and E. Hirota, J. Chem. Phys., M. J. McClements and G. A. Takacs, J. Phys. Chem., 1978,
1988, 89, 621–624. 82, 2252–2257.
102 A. Rakhymzhan and A. Chichinin, J. Phys. Chem. A, 2010, 107 W. M. Gelbart, Annu. Rev. Phys. Chem., 1977, 28,
114, 6586–6593. 323–348.
Published on 21 February 2014. Downloaded by Academia Sinica - Taipei on 15/12/2017 09:16:31.

103 M. Roth, C. Maul and K.-H. Gericke, Phys. Chem. Chem. 108 J. M. Bowman and B. C. Shepler, Annu. Rev. Phys. Chem.,
Phys., 2002, 4, 2932–2940. 2011, 62, 531–553.

7198 | Phys. Chem. Chem. Phys., 2014, 16, 7184--7198 This journal is © the Owner Societies 2014

Вам также может понравиться