Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Article

pubs.acs.org/JPCA

Gas-Phase Generation and Matrix Isolation of the Methylsulfonyl


Radical CH3SO2• from Allylmethylsulfone
Hans Peter Reisenauer,*,† Peter R. Schreiner,† Jaroslaw Romanski,‡ and Grzegorz Mloston*,‡

Institute of Organic Chemistry, Justus-Liebig University, Heinrich-Buff-Ring 58, D-35392 Giessen, Germany

Department of Organic and Applied Chemistry, University of Lodz, Tamka 12, 91-493 Lodz, Poland
*
S Supporting Information

ABSTRACT: The atmospherically highly relevant methyl-


sulfonyl radical (CH3SO2•) was generated by high-vacuum
flash pyrolysis (HVFP) of allylmethylsulfone and isolated in an
argon matrix at 10 K; the allyl radical formed as the
cofragment. Upon thermolysis, the methylsulfonyl radical
undergoes partial decomposition, leading to substantial
amounts of sulfur dioxide in the matrix. The title compound
was characterized through the assignment of eight fundamental
IR bands of its CD3 and 13CH3 isotopologues and the excellent
agreement with the B3LYP/6-311+G(3df,3pd) computed
harmonic vibrational frequencies. The two most intense
absorptions were found at 1267.1 and 1067.6 cm−1. In
extension of this study S-methyl methanethiosulfonate was found to be another suitable, although less efficient, precursor for the
gas-phase generation of the methylsulfonyl radical.

■ INTRODUCTION
Sulfur-containing radicals derived from oxidation processes of
mixtures.10 The IR frequencies of 2, generated through the
reaction of methyl radicals and sulfur dioxide, were reported for
volatile sulfur organic compounds in the atmosphere have the gas phase and for 2 isolated in para-hydrogen matrixes;
extensively been investigated because of their relevance in the these compare favorably with computed values.11,12 Radical 2
multiple reaction pathways eventually leading to atmospheric was also identified as a short-lived intermediate in the pulse
sulfate aerosols. These affect the earth’s radiation balance either radiolysis of methanesulfonyl chloride through its UV
by direct backscattering of solar radiation or indirectly via the absorption at λmax ≅ 320 nm.6−8,13 The structure and reactivity
cloud albedo by forming cloud condensation nuclei.1−5 Two of 2 were the subject of many theoretical studies at the density
distinct representatives of this class of radicals species are the functional theory (DFT) and ab initio (up to CCSD(T)/F12)
methylsulfinyl (1) and methylsulfonyl radicals (2) (Scheme 1). levels.11,12,14−17
In a recent paper, we reported an efficient method for the
Scheme 1. Methylsulfinyl (1) and Methylsulfonyl Radicals generation and matrix isolation of 1 by high-vacuum flash
(2) pyrolysis (HVFP) of allylmethylsulfide S-oxide (3).18 This
success prompted us to utilize an analogous protocol for the
generation of 2 from allylmethylsulfone (4) (Scheme 2).
The specific challenge of this strategy lies in the fact that the
C−S bond in 2 is very weak and the dissociation leading to the
methyl radical and sulfur dioxide can easily occur (Scheme 2).
The formation of 1 and 2 in the atmosphere is believed to Experimental and computational (B3LYP/aug-cc-pVTZ) esti-
occur via oxidation of dimethyl sulfide produced in large mates provide a C-S bond dissociation energy (BDE) of only
amounts by biodegradation of phytoplankton present in the about 14 kcal mol−1.12,16 Therefore, the subsequent dissociation
oceans.1−5 of the target radical 2 into sulfur dioxide and the methyl radical
Sulfonyl radicals (R−SO2•) are also widely applied as is expected to proceed readily under the pyrolysis conditions,
versatile intermediates in organic synthesis, for example, for whereas the comparable cleavage of 1 into the methyl radical
the preparation of unsaturated sulfones,6−8 and as universal
leaving groups in reversible addition−fragmentation chain- Special Issue: Markku Räsänen Festschrift
transfer polymerization.9 Particular attention focuses on 2, and
its spectroscopic properties were described in recently Received: April 14, 2014
published reports.6−8,10−17 ESR spectroscopy allowed the Revised: May 24, 2014
detection of 2 as a radiolysis product of H2O/DMSO Published: May 28, 2014

© 2014 American Chemical Society 2211 dx.doi.org/10.1021/jp5036647 | J. Phys. Chem. A 2015, 119, 2211−2216
The Journal of Physical Chemistry A Article

Scheme 2. B3LYP/6-311+G(3df,3pd) Assessment of the (4): −6 °C; S-methyl methanethiosulfonate (5): −25 °C) into
Bond Dissociation Energies (BDEs) for the Formation of the pyrolysis tube. Immediately after leaving the tube, at a
the Methylsulfinyl (1) and Methylsulfonyl (2) Radicalsa distance of ∼50 mm, the pyrolysis products were co-condensed
with a large excess of argon on the surface of the 10 K cold
matrix window during a time of 2−4 h. Pyrolysis temperatures
were varied between 500 and 900 °C; the optimal temperatures
with respect to the yields of methylsulfonyl radicals (2) were
800 (4) and 850 °C (3). For irradiations, a mercury high-
pressure lamp (Osram, HBO 200 in connection with a Bausch
& Lomb monochromator; slits set to a spectral half-width of
Δλ = 10 nm) was used. An unspecific photodegradation of 2 to
COS, CO, H2S, and H2O was observed with λ = 313 nm (7 h
a
Values in kcal mol−1.
for complete conversion). Subtraction of the corresponding
spectra before and after photodegradation was made to extract
and SO requires much more energy (about 23 kcal mol−1) and the IR and UV/vis spectra of 2. The same procedure was
is thus very unlikely to occur. Moreover, our B3LYP/6- applied for the CD3 and 13C isotopologues.
311+G(3df,3pd) computations predict that the energy required
for splitting off the allyl radical from the S,S-dioxide 4 is 16.2
kcal mol−1 higher than that for the same process for S-
■ COMPUTATIONAL METHODS
We utilized DFT to assess reaction energies and equilibrium
monoxide 3.


geometries, harmonic vibrational wavenumbers, as well as IR
intensities of all key structures using the B3LYP23,24 functional
EXPERIMENTAL SECTION with a large basis set including polarization and diffuse
Preparation of the Precursor. The preparation of 4 and functions (6-311+G(3df,3pd)) as implemented in the GAUS-
its isotopologues was based on known proctocols.19−22 A SIAN 09 program.25 All stationary structures were characterized
solution of ally1 bromide (1.94 g, 0.016 mol) and thiourea (1.0 as minima or transition structures by computing analytical
g, 0.016 mol) in methanol (20 mL) was refluxed for 1 h, and harmonic vibrational frequencies (cf. the Suppporting In-
then, after addition of a methanolic solution of NaOH (1.28 g formation (SI)).
in 15 mL), heating was continued for another hour. Next, the
solution was cooled to −10 °C, and MeI (2.28 g, 0.016 mol)
was added in small portions. The mixture obtained thereby was
■ RESULTS AND DISCUSSION
Allylmethylsulfone (4) can readily be prepared through the
refluxed for the next 0.5 h. After this time, the mixture was sequence of reactions depicted in Scheme 3.19−22
distilled, and the allylmethylsulfide was collected as an
azeotrope with methanol (bp 62−64 °C). A solution of Scheme 3. Two-Step Synthesis of Allylmethylsulfone (4)
allylmethylsulfide (1.0 g, 11.0 mmol) in glacial acetic acid (3
cm3) was chilled in ice, and the 30% aq hydrogen peroxide (2
or 3 times excess) was added dropwise, and the mixture was
stirred overnight. The solution was heated at 85 °C for 1 h,
then cooled, and diluted with an equal volume of water. The
aqueous solution was extracted twice with methylene chloride.
The combined organic layers were washed with water and dried The parent compound and its CD3 and 13CH3 isotopologues
over anhydrous MgSO4. The solvents were evaporated under were prepared by treatment of sodium allylthiolate with CH3I,
reduced pressure, and the residue was distilled at 120 °C/0.1 CD3I, and 13CH3I, respectively, and were oxidized with an
mm Hg (Kugelrohr) to yield 0.8 g of product. Because traces of access of H2O2 in glacial acetic acid. Distillation yielded
allylmethylsulfoxide could not be removed by distillation, colorless oils that were used for HVFP experiments. Reaction
preparative GC was applied to get the necessary amounts of conditions were optimized with respect to a maximum yield of
pure sulfone 4 for the matrix experiments. Commercially 2 by several test runs and eventually established 800 °C at
available S-methyl methanethiosulfonate (5) was used without ∼10−5 mbar as the optimal temperature and pressure. The
further purification. pyrolysate was condensed with a large excess of argon at 10 K
Matrix Isolation Experiments. The cryostat used for the on the surface of the spectroscopic window to form a solid
matrix isolation studies was an APD Cryogenics HC-2 closed- matrix, which subsequently was analyzed by means of IR and
cycle refrigerator system fitted with CsI windows for IR and UV/vis spectroscopy. The IR spectrum (Figure 1) revealed the
BaF2 windows for UV/vis measurements. IR spectra were presence of strong absorption bands at 1535/1531, 1151/1147,
recorded with a Bruker IFS 55 FTIR spectrometer (4500−300 and 519/517 cm−1, which could easily be attributed to sulfur
cm−1; resolution: 0.7 cm−1), and UV/vis spectra were recorded dioxide (SO2) by comparison with an argon matrix spectrum of
with a JASCO V-670 spectrophotometer (spectral range: 190− a pure sample.
2500 nm; resolution: ≥ 0.1 nm). We employed a small home- Another set of absorption bands located at 801, 984, 1389,
built water-cooled oven directly connected to the vacuum and 1477 cm−1 provides evidence for the presence of the allyl
shroud of the cryostat for the combination of HVFP with radical formed as a coproduct of the thermal decomposition of
matrix isolation. The pyrolysis zone consisted of an empty 4.26−29 Finally, the presence of a set of narrow bands of
quartz tube (inner diameter, 8 mm; length of heating zone, 50 medium intensity at 1413.8, 1267.1, 1067.6, 631.3, 460.1, and
mm) resistively heated by a coax heating wire; the temperature 385.5 cm−1 shows the formation of 2 (Figure 2, Table 1). The
was controlled by a Ni/CrNi thermocouple. The precursors first four values correspond well with the recently reported IR
were evaporated from a cooled storage bulb (allylmethylsulfone data of matrix-isolated 2 in solid para-hydrogen.11 The overall
2212 dx.doi.org/10.1021/jp5036647 | J. Phys. Chem. A 2015, 119, 2211−2216
The Journal of Physical Chemistry A Article

Figure 1. FT-IR spectrum of the matrix-isolated pyrolysis products of allylmethylsulfone (4) (Ar, 10 K; pyrolysis temperature: 800 °C). Bands
attributed to the allyl radical are marked with an asterisk (∗).

Figure 2. (A) Computed (B3LYP/6-311+G(3df,3pd)) and (B) experimental (matrix-isolated, Ar, 10 K, bottom) IR spectrum of the CH3SO2•
radical. The experimental spectrum is a difference of the spectra taken before and after 5 h of irradiation with 313 nm. Bands attributed to the allyl
radical are marked with asterisks (∗).

observed spectral pattern fits nicely to the DFT-computed IR expected products. However, as apparent from the spectrum in
spectrum of 2. With the aid of the computations, the additional Figure 1, a significant amount of unconverted 4 was also
two weak low-frequency bands found at 460.1 and 385.5 cm−1 present. This observation confirms the computational pre-
could be assigned to the (OSO) bending and to the SO2 diction that the BDE for the C−S bond in 4 is 16 kcal mol−1
wagging mode, respectively (Figure 2). Moreover, the
higher than that of 3.
experimentally observed spectral band shifts upon deuteration
and 13C substitution of the methyl group agree well with the Note that apart from the allyl and methylsulfonyl radicals,
values computed within the harmonic approximation (Table 1). also methyl radicals and sulfur dioxide (SO2) formed apparently
The presence of SO2, methyl radical, allyl radical, and 2 via secondary fragmentation of 2 in the gas phase. According to
demonstrates that the thermal decomposition of 4 led to the our computations, the S−C BDE of 2 only amounts to 14 kcal
2213 dx.doi.org/10.1021/jp5036647 | J. Phys. Chem. A 2015, 119, 2211−2216
The Journal of Physical Chemistry A Article

Table 1. Computed and Experimental IR Spectra of CH3SO2• (2) and Band Shifts for the 13CH3SO2 (13C-2), and CD3SO2• (d3-
2) Radicalsa
CH3SO2• (2) 13
CH3SO2• (13C-2) CD3SO2• (d3-2)
−1 −1
band position [cm ] band shifts [cm ] band shifts [cm−1]
sym. approx. mode comput. expt. comput. expt. comput. expt.
ν10, a″ CH str. 3185.1 (0.0) n.o. −12.2 n.o. −819.4 n.o.
ν1, a′ CH str. 3164.2 (0.5) n.o. −11.8 n.o. −817.1 n.o.
ν2, a′ CH str. 3060.6 (0.2) n.o. −12.4 n.o. −873.2 n.o.
ν3, a′ CH3 def. 1455.3 (9.7) 1413.9 (w) −2.6 −2.3 −402.9 −382.5
ν11, a″ CH3 def. 1446.8 (5.1) n.o. −2.5 n.o. −401.0 n.o.
ν4, a′ CH3 def. 1312.3 (0.9) n.o. −7.3 n.o. −312.0 n.o.
ν12, a″ SO2 str. 1271.1 (141.4) (34S: 1254.7) 1267.1 (vs) (34S: 1251.2) 0.0 0.0 −0.4 +0.3
ν5, a′ SO2 str. 1076.4 (66.1) (34S: 1069.2) 1067.6 (s) (34S: 1060.7) 0.0 −3.2 +1.2 +5.1
ν13, a″ CH3 rock. 955.5 (0.0) n.o. −7.1 n.o. −214.7 n.o.
ν6, a′ CH3 rock. 933.8 (6.8) 915.8 (w) −6.1 −5.7 −200.9 −189.0
ν7, a′ CS str. 624.8 (16.2) 631.4 (w) −12.1 −11.4 −28.3 −28.1
ν8, a′ SO2 def. 459.2 (15.7) 460.1 (w) −1.2 −1.1 −9.4 −9.0
ν9, a′ SO2 wag. 380.9 (21.3) 385.5 (w) −2.6 −2.1 −23.6 −25.2
ν14, a″ SO2 rock. 302.1 (0.1) n.o. −3.7 n.o. −26.1 n.o.
ν15, a″ CH3 twist. 173.5 (0.1) n.o. 0.0 n.o. −47.1 n.o.
a
Computations: UB3LYP/6-311+G(3df,3pd), harmonic approximation, unscaled frequencies; experiment: argon matrix, 10 K. Intensities (in
parentheses): in km mol−1 for the computations; approximate relative intensities (w: weak, s: strong, vs: very strong) of the experiments; n.o. = not
observed.

mol−1, and dissociation is expected in the gas phase before spectral region between 300 and 450 nm disappeared, while the
trapping in the argon matrix. characteristic absorption bands of the allyl radical were
The UV/vis spectra of the matrix-isolated pyrolysate exhibit unchanged (Figure 3). A difference spectrum taken before
the known absorption bands of the allyl radical,26 which overlap and after photolysis shows a broad absorption band with a
with additional broad absorptions located between 300 and 450 maximum at around 340 nm (cf. SI) that is very similar to the
nm and originating from SO2 and 2 (Figure 3). Upon spectrum of 2 in solution.6−8,13
Along with the UV/vis spectra, a corresponding IR analysis
was performed to monitor the course of photolysis. A mixture
of COS, CO, and H2O was identified based on their specific IR
absorption bands. Although the photolysis of 2 led to a rather
unspecific decomposition, it proved to be very helpful in its IR
analysis through calculating meaningful difference spectra
(Figure 2).
Another precursor of 2 tested in the present study was S-
methyl methanethiosulfonate (5), which was expected to give 2
along with the mercaptomethyl radical (CH3S•) (6)30 via
homolytic cleavage of the S−S bond, although concerted
elimination of thioformaldehyde yielding methanesulfinic acid
(7) may compete energetically (Scheme 4). For this process,
we computed an activation enthalpy of 38.7 kcal mol−1,
whereas the dissociation energy of the S−S bond was 45.0 kcal
mol−1.
In a HVFP experiment performed at 750 °C, almost no
Figure 3. UV/vis spectra of the matrix-isolated pyrolysate of decomposition of 5 was observed. Even if the temperature of
allylmethylsulfone (3). Solid line: not irradiated; dashed line: after 5
h of irradiation with 313 nm light.
the hot zone was increased to 850 °C, the product mixture in
the argon matrix still contained starting material as the major
component (cf. Figure S1, SI). On the other hand, absorption
photolysis using 312 nm light (mercury high-pressure lamp bands at 1413.8, 1267.1, 1067.6, 631.3, and 460.1 cm−1
with a monochromator), the broad absorption in the UV confirmed the presence of 2. In addition, the bands located

Scheme 4. Competing Thermal Decomposition Pathways of S-Methylmethanethiosulfonate (5)

2214 dx.doi.org/10.1021/jp5036647 | J. Phys. Chem. A 2015, 119, 2211−2216


The Journal of Physical Chemistry A Article

at 987.4 and 1055.2 cm−1 provided evidence for the formation (6) Gilmore, K.; Gold, B.; Clark, R. J.; Alabugin, I. V. Convenient
of thioformaldehyde (CH2S).30,31 Similarly to the experiment Ambient Temperature Generation of Sulfonyl Radicals. Aust. J. Chem.
carried out with 4, the pyrolysate contained substantial 2013, 66, 336−340.
amounts of sulfur dioxide (SO2). (7) Chatgilialoglu, C.; Mozziconacci, O.; Tamba, M.; Bobrowski, K.;


Kciuk, G.; Bertrand, M. P.; Gastaldi, S.; Timokhin, V. I. Kinetic Studies
on the Formation of Sulfonyl Radicals and Their Addition to Carbon−
CONCLUSION Carbon Multiple Bonds. J. Phys. Chem. A 2012, 116, 7623−7628.
The present study shows that the methylsulfonyl radical (2) is (8) Tamba, M.; Dajka, K.; Ferreri, C.; Asmus, K.-D.; Chatgilialoglu,
kinetically much less stable than the previously reported C. One-Electron Reduction of Methanesulfonyl Chloride. The Fate of
methylsulfinyl radical (1). The reason for the observed MeSO2Cl•− and MeSO2• Intermediates in Oxygenated Solutions and
differences in the stability of both radicals is the high Their Role in the Cis−Trans Isomerization of Mono-unsaturated Fatty
Acids. J. Am. Chem. Soc. 2007, 129, 8716−8723.
dissociation energy of the C−S bond in 1 computed to be (9) Gryn’ova, G.; Guliashvili, T.; Matyjaszewski, K.; Coote, M. L.
47.5 kcal mol−1. In marked contrast, the BDE of the Computational Evaluation of the Sulfonyl Radical as a Universal
corresponding bond in 2 is computed to be only 14.6 kcal Leaving Group for RAFT Polymerisation. Aust. J. Chem. 2013, 66,
mol−1. Despite these unfavorable energetics, HVFP can be used 308−313.
to generate 2 using readily available allylmethylsulfone (4) as a (10) Dondi, D.; Buttafava, A.; Zeffiro, A.; Cherubini, R.; De Nadal,
precursor. This method establishes a new convenient approach V.; Gerardi, S.; Faucitano, A. The Origin of the Radiobiological
for the gas-phase generation of 2 based on a single precursor Damage in Cells Stored in Cryostatic Conditions. Radiat. Phys. Chem.
molecule. 2012, 81, 1445−1450.


(11) Lee, Y.-F.; Lee, Y.-P. Infrared Absorption of CH3SO2 Observed
upon Irradiation of a p-H2 Matrix Containing CH3I and SO2. J. Chem.
ASSOCIATED CONTENT Phys. 2011, 134, 124314.
*
S Supporting Information (12) Chu, L.-K.; Lee, Y.-P. Infrared Absorption of CH3SO2 Detected
Additional spectroscopic data and computational results with Time-Resolved Fourier-Transform Spectroscopy. J. Chem. Phys.
(optimized geometry, energies, Cartesian coordinates, and 2006, 124, 244301.
vibrational harmonic frequencies of optimized geometries). (13) Chatgilialoglu, C.; Griller, D.; Guerra, M. Experimental and
Theoretical Approaches to the Optical Absorption Spectra of Sulfonyl
This material is available free of charge via the Internet at
Radicals. J. Phys. Chem. 1987, 91, 3747−3750.
http://pubs.acs.org.


(14) Jørgensen, S.; Jensen, C.; Kjaergaard, H. G.; Anglada, J. M. The
Gas-Phase Reaction of Methane Sulfonic Acid with the Hydroxyl
AUTHOR INFORMATION Radical without and with Water Vapor. Phys. Chem. Chem. Phys. 2013,
Corresponding Authors 15, 5140−5150.
(15) Salta, Z.; Kosmas, A. M.; Lesar, A. D. Computational
*E-mail: Hans.P.Reisenauer@org.chemie.uni-giessen.de. Tele- Investigation of the Peroxy Radicals CH3S(O)nOO and the
phone: (+49)-6419934380 (H.P.R.). Peroxynitrates CH3S(O)nOONO2 (n=0, 1, 2). Comput. Theor. Chem.
*E-mail: gmloston@uni.lodz.pl (G.M.). 2012, 1001, 67−76.
Author Contributions (16) Mandal, D.; Bagchi, S.; Das, A. K. Mechanism and Kinetics for
The ideas were conceived and the experiments planned by all the Reaction of O(3P) with DMSO: A Theoretical Study. Chem. Phys.
authors. Measurements and computations were carried out by Lett. 2012, 551, 31−37.
(17) Ratliff, B. J.; Tang, X.; Butler, L. J.; Szpunar, D. E.; Lau, K.-C.
H.P.R. All authors wrote the manuscript. All authors have given Determining the CH3SO2→CH3+SO2 Barrier from Methylsulfonyl
approval to the final version of the manuscript. Chloride Photodissociation at 193 nm Using Velocity Map Imaging. J.
Notes Chem. Phys. 2009, 131, 044304.
The authors declare no competing financial interest. (18) Reisenauer, H. P.; Romanski, J.; Mloston, G.; Schreiner, P. R.


Matrix Isolation and Spectroscopic Properties of the Methylsulfinyl
Radical CH3(O)S•. Chem. Commun. 2013, 49, 9467−9469.
ACKNOWLEDGMENTS (19) Prabhuswamy, B.; Ambekar, S. Y. Synthesis of 3a,4-Dihydro-8-
This work was supported by the DAAD (University Partnership substituted-3H-isoxazolo[c-4,3]thiapyrano[5,6-3,2]quinolines. Synth.
between the University of Lodz and the Justus-Liebig Commun. 1999, 29, 3477−3485.
University). G.M. acknowledges the National Science Center (20) Kline, M. L.; Beutow, N.; Kim, J. K.; Caserio, M. C.
(PL-Cracow) for financial support (Grant Maestro−3 (Dec− Methylthiolation of Unsaturated Sulfides. Thiosulfonium Ions. J. Org.
2012/06/A/ST5/00219). Chem. 1979, 44, 1904−1910.


(21) O’Connor, D. E.; Lyness, W. I. The Effect of Methylmercapto,
Methylsulfinyl, and Methylsulfonyl Groups on the Equilibrium in
REFERENCES Three-Carbon Prototropic Systems. J. Am. Chem. Soc. 1964, 86, 3840−
(1) Charlson, R. J.; Lovelock, J. E.; Andreae, M. O.; Warren, S. G. 3846.
Oceanic Phytoplankton, Atmospheric Sulphur, Cloud Albedo and (22) Culshaw, P. N.; Walton, J. C. Sulphonate Esters as Sources of
Climate. Nature 1987, 326, 655−661. Sulphonyl Radicals; Ring-Closure Reactions of Alk-4- and -5-
(2) Albrecht, B. Aerosols, Cloud Microphysics, and Fractional Enesulphonyl Radicals. J. Chem. Soc., Perkin Trans. 2 1991, 1201−
Cloudiness. Science 1989, 245, 1227−1230. 1208.
(3) Charlson, R. J.; Schwarz, S. E.; Hales, J. M.; Cess, R. D.; Coakley, (23) Becke, A. D. Density-Functional Exchange-Energy Approx-
J. A., Jr.; Hansen, J. E.; Hofmann, D. J. Climate Forcing by imation with Correct Asymptotic Behavior. Phys. Rev. A 1988, 38,
Anthropogenic Aerosols. Science 1992, 255, 423−430. 3098−3100.
(4) Andreae, M. O.; Jones, C. D.; Cox, P. M. Strong Present-Day (24) Miehlich, B.; Savin, A.; Stoll, H.; Preuss, H. Results Obtained
Aerosol Cooling Implies a Hot Future. Nature 2005, 435, 1187−1190. with the Correlation-Energy Density Functionals of Becke and Lee,
(5) Faloona, I. Sulfur Processing in the Marine Atmospheric Yang and Parr. Chem. Phys. Lett. 1989, 157, 200−206.
Boundary Layer: A Review and Critical Assessment of Modeling (25) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.;
Uncertainties. Atmos. Environ. 2009, 43, 2841−2854. Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci,

2215 dx.doi.org/10.1021/jp5036647 | J. Phys. Chem. A 2015, 119, 2211−2216


The Journal of Physical Chemistry A Article

B.; Petersson, G. A.; et al. Gaussian09, revision D.01; Gaussian Inc.:


Wallingford, CT, 2013.
(26) Maier, G.; Reisenauer, H. P.; Rohde, B.; Dehnicke, K. IR-, UV-
und ESR-Spektrum des Allylradikals in einer Argon-Matrix. Chem. Ber.
1983, 116, 732−740.
(27) Mal’tsev, A. K.; Korolov, V. A.; Nefedov, O. M. First Direct
Infrared Study of the Free Allyl Radical. Bull. Acad. Sci. USSR, Div.
Chem. Sci. 1982, 2131−2131.
(28) Mal’tsev, A. K.; Korolov, V. A.; Nefedov, O. M. IR Spectroscopic
Study of Matrix-Isolated Free Allyl Radicals and Analysis of Vibrational
Spectra and Structure of π-Allyl Organometallic Compounds. Bull.
Acad. Sci. USSR, Div. Chem. Sci. 1984, 510−521.
(29) Holtzhauer, K.; Cometta-Morini, C.; Oth, J. F. M. Photo-
chemical Electrocyclization of the Allyl Radical into the Cyclopropyl
Radical. J. Phys. Org. Chem. 1990, 219−229.
(30) Jacox, M. E.; Milligan, D. E. Matrix Isolation Study of the
Infrared Spectrum of Thioformaldehyde. J. Mol. Spectrosc. 1975, 58,
142−157.
(31) Torres, M.; Safarik, I.; Clement, I. A.; Strausz, O. P. The
Generation and Vibrational Spectrum of Matrix Isolated Thioformal-
dehyde and Dideuterothioformaldehyde. Can. J. Chem. 1982, 60,
1187−1191.

2216 dx.doi.org/10.1021/jp5036647 | J. Phys. Chem. A 2015, 119, 2211−2216

You might also like