4 research outputs found

    Detection and Characterization of Products from Photodissociation of XCH<sub>2</sub>CH<sub>2</sub>ONO (X = F, Cl, Br, OH)

    No full text
    Alkyl nitrites have been used previously to produce alkoxy radicals, which are important intermediates in the oxidation of alkanes in atmospheric and combustion processes. Substituted alkoxy radicals, particulary hydroxyalkoxy radicals, are also important intermediates in the atmospheric oxidation of alkenes and combustion of alcohols. In order to produce substituted alkoxy radicals we have photolyzed at 351 nm substituted alkyl nitrites, XCH<sub>2</sub>CH<sub>2</sub>ONO (X = F, Cl, Br, OH). Using laser-induced fluorescence only in the case of X = F do we observe the spectrum of substituted alkoxy radical, XCH<sub>2</sub>CH<sub>2</sub>O; but we always observe the electronic transitions of formaldehyde, HCHO, and vinoxy radical, CH<sub>2</sub>CHO. HCHO can be formed by the dissociation of XCH<sub>2</sub>CH<sub>2</sub>O in its ground state as the barrier to C–C bond dissociation is less than the photon energy remaining after O–NO bond breakage. However, the barrier along the reaction path directly leading from XCH<sub>2</sub>CH<sub>2</sub>O to CH<sub>2</sub>CHO + HX is much higher than the available energy remaining after O–NO bond breakage. A roaming mechanism, involving a frustrated dissociation of X followed by HX extraction, might explain the apparent paradox. Under the conditions of our observations vinoxy retains considerable vibrational excitation but the observed rotational temperatures of both HCHO and CH<sub>2</sub>CHO are ≲7 K

    Observation of the Ã−X̃ Electronic Transition of the β-Hydroxyethylperoxy Radical

    No full text
    The Ã−X̃ electronic absorption spectrum of β-hydroxyethyl peroxy radical (β-HEP) has been recorded in the NIR by cavity ringdown spectroscopy. The precursor, β-hydroxyethyl radical, is generated by photolysis of 2-iodoethanol and by OH-initiated oxidation of ethene, in both cases followed by addition of O<sub>2</sub> to form the peroxy. Although electronic structure calculations predict that 13 conformers of β-HEP exist as minima on the potential energy surface, the experimental spectrum is rationalized in terms of the band origins and vibrational progressions of only the two most stable conformers, G<sub>1</sub>G<sub>2</sub>G<sub>3</sub> and G<sub>1</sub><sup>′</sup>G<sub>2</sub>G<sub>3</sub>

    Imaging and Scattering Studies of the Unimolecular Dissociation of the BrCH<sub>2</sub>CH<sub>2</sub>O Radical from BrCH<sub>2</sub>CH<sub>2</sub>ONO Photolysis at 351 nm

    No full text
    We report a study of the unimolecular dissociation of BrCH<sub>2</sub>CH<sub>2</sub>O radicals produced from the photodissociation of BrCH<sub>2</sub>CH<sub>2</sub>ONO at 351/355 nm. Using both a crossed laser-molecular beam scattering apparatus with electron bombardment detection and a velocity map imaging apparatus with tunable VUV photoionization detection, we investigate the initial photodissociation channels of the BrCH<sub>2</sub>CH<sub>2</sub>ONO precursor and the subsequent dissociation of the vibrationally excited BrCH<sub>2</sub>CH<sub>2</sub>O radicals. The only photodissociation channel of the precursor we detected upon photodissociation at 351 nm was O–NO bond fission. C–Br photofission and HBr photoelimination do not compete significantly with O–NO photofission at this excitation wavelength. The measured O–NO photofission recoil kinetic energy distribution peaks near 14 kcal/mol and extends from 5 to 24 kcal/mol. There is also a small signal from lower kinetic energy NO product (it would be 6% of the total if it were also from O–NO photofission). We use the O–NO photofission <i>P</i>(<i>E</i><sub>T</sub>) peaking near 14 kcal/mol to help characterize the internal energy distribution in the nascent ground electronic state BrCH<sub>2</sub>CH<sub>2</sub>O radicals. At 351 nm, some but not all of the BrCH<sub>2</sub>CH<sub>2</sub>O radicals are formed with enough internal energy to unimolecularly dissociate to CH<sub>2</sub>Br + H<sub>2</sub>CO. Although the signal at <i>m</i>/<i>e</i> = 93 (CH<sub>2</sub>Br<sup>+</sup>) obtained with electron bombardment detection includes signal both from the CH<sub>2</sub>Br product and from dissociative ionization of the energetically stable BrCH<sub>2</sub>CH<sub>2</sub>O radicals, we were able to isolate the signal from CH<sub>2</sub>Br product alone using tunable VUV photoionization detection at 8.78 eV. We also sought to investigate the source of vinoxy radicals detected in spectroscopic experiments by Miller and co-workers (J. Phys. Chem. A 2012, 116, 12032) from the photodissociation of BrCH<sub>2</sub>CH<sub>2</sub>ONO at 351 nm. Using velocity map imaging and photodissociating the precursor at 355 nm, we detected a tiny signal at <i>m</i>/<i>e</i> = 43 and a larger signal at <i>m</i>/<i>e</i> = 15 that we tentatively assign to vinoxy. An underlying signal in the time-of-flight spectra at <i>m</i>/<i>e</i> = 29 and <i>m</i>/<i>e</i> = 42, the two strongest peaks in the literature electron bombardment mass spectrum of vinoxy, is also apparent. Comparison of those signal strengths with the signal at HBr<sup>+</sup>, however, shows that the vinoxy product does not have HBr as a cofragment, so the prior suggestion by Miller and co-workers that the vinoxy might result from a roaming mechanism is contraindicated

    Direct Measurements of Unimolecular and Bimolecular Reaction Kinetics of the Criegee Intermediate (CH<sub>3</sub>)<sub>2</sub>COO

    No full text
    The Criegee intermediate acetone oxide, (CH<sub>3</sub>)<sub>2</sub>COO, is formed by laser photolysis of 2,2-diiodopropane in the presence of O<sub>2</sub> and characterized by synchrotron photoionization mass spectrometry and by cavity ring-down ultraviolet absorption spectroscopy. The rate coefficient of the reaction of the Criegee intermediate with SO<sub>2</sub> was measured using photoionization mass spectrometry and pseudo-first-order methods to be (7.3 ± 0.5) × 10<sup>–11</sup> cm<sup>3</sup> s<sup>–1</sup> at 298 K and 4 Torr and (1.5 ± 0.5) × 10<sup>–10</sup> cm<sup>3</sup> s<sup>–1</sup> at 298 K and 10 Torr (He buffer). These values are similar to directly measured rate coefficients of <i>anti</i>-CH<sub>3</sub>CHOO with SO<sub>2</sub>, and in good agreement with recent UV absorption measurements. The measurement of this reaction at 293 K and slightly higher pressures (between 10 and 100 Torr) in N<sub>2</sub> from cavity ring-down decay of the ultraviolet absorption of (CH<sub>3</sub>)<sub>2</sub>COO yielded even larger rate coefficients, in the range (1.84 ± 0.12) × 10<sup>–10</sup> to (2.29 ± 0.08) × 10<sup>–10</sup> cm<sup>3</sup> s<sup>–1</sup>. Photoionization mass spectrometry measurements with deuterated acetone oxide at 4 Torr show an inverse deuterium kinetic isotope effect, <i>k</i><sub>H</sub>/<i>k</i><sub>D</sub> = (0.53 ± 0.06), for reactions with SO<sub>2</sub>, which may be consistent with recent suggestions that the formation of an association complex affects the rate coefficient. The reaction of (CD<sub>3</sub>)<sub>2</sub>COO with NO<sub>2</sub> has a rate coefficient at 298 K and 4 Torr of (2.1 ± 0.5) × 10<sup>–12</sup> cm<sup>3</sup> s<sup>–1</sup> (measured with photoionization mass spectrometry), again similar to rate for the reaction of <i>anti</i>-CH<sub>3</sub>CHOO with NO<sub>2</sub>. Cavity ring-down measurements of the acetone oxide removal without added reagents display a combination of first- and second-order decay kinetics, which can be deconvolved to derive values for both the self-reaction of (CH<sub>3</sub>)<sub>2</sub>COO and its unimolecular thermal decay. The inferred unimolecular decay rate coefficient at 293 K, (305 ± 70) s<sup>–1</sup>, is similar to determinations from ozonolysis. The present measurements confirm the large rate coefficient for reaction of (CH<sub>3</sub>)<sub>2</sub>COO with SO<sub>2</sub> and the small rate coefficient for its reaction with water. Product measurements of the reactions of (CH<sub>3</sub>)<sub>2</sub>COO with NO<sub>2</sub> and with SO<sub>2</sub> suggest that these reactions may facilitate isomerization to 2-hydroperoxypropene, possibly by subsequent reactions of association products
    corecore