Solubility and solution-phase chemistry of isocyanic acid, methyl isocyanate, and cyanogen halides

<p>Condensed-phase uptake and reaction are important atmospheric removal processes for reduced nitrogen species, isocyanic acid (HNCO), methyl isocyanate (<span class="inline-formula">CH<sub>3</sub>NCO</span>), and cyanogen halides (XCN, X&thinsp;<span...

Full description

Bibliographic Details
Main Authors: J. M. Roberts, Y. Liu
Format: Article
Language:English
Published: Copernicus Publications 2019-04-01
Series:Atmospheric Chemistry and Physics
Online Access:https://www.atmos-chem-phys.net/19/4419/2019/acp-19-4419-2019.pdf
Description
Summary:<p>Condensed-phase uptake and reaction are important atmospheric removal processes for reduced nitrogen species, isocyanic acid (HNCO), methyl isocyanate (<span class="inline-formula">CH<sub>3</sub>NCO</span>), and cyanogen halides (XCN, X&thinsp;<span class="inline-formula">=</span>&thinsp;Cl, Br, I); yet many of the fundamental quantities that govern this chemistry have not been measured or are not well studied. These nitrogen species are of emerging interest in the atmosphere as they have either biomass burning sources, i.e., HNCO and <span class="inline-formula">CH<sub>3</sub>NCO</span>, or, like the XCN species, have the potential to be a significant condensed-phase source of <span class="inline-formula">NCO<sup>−</sup></span> and therefore HNCO. Solubilities and the first-order reaction rate of these species were measured for a variety of solutions using a bubble flow reactor method with total reactive nitrogen (<span class="inline-formula">N<sub>r</sub></span>) detection. The aqueous solubility of HNCO was measured as a function of pH and had an intrinsic Henry's law solubility of 20 (<span class="inline-formula">±2</span>)&thinsp;M&thinsp;atm<span class="inline-formula"><sup>−1</sup></span> and a <span class="inline-formula"><i>K</i><sub>a</sub></span> of 2.0 (<span class="inline-formula">±0.3</span>)&thinsp;<span class="inline-formula">×</span>&thinsp;10<span class="inline-formula"><sup>−4</sup></span>&thinsp;M (<span class="inline-formula"><i>p</i><i>K</i><sub>a</sub></span>&thinsp;<span class="inline-formula">=</span>&thinsp;<span class="inline-formula">3.7±0.1</span>) at 298&thinsp;K. The temperature dependence of HNCO solubility was very similar to other small nitrogen-containing compounds, such as HCN, acetonitrile (<span class="inline-formula">CH<sub>3</sub>CN</span>), and nitromethane, and the dependence on salt concentration exhibited the “salting out” phenomenon. The rate constant of reaction of HNCO with 0.45&thinsp;M <span class="inline-formula"><math xmlns="http://www.w3.org/1998/Math/MathML" id="M16" display="inline" overflow="scroll" dspmath="mathml"><mrow class="chem"><msubsup><mi mathvariant="normal">NH</mi><mn mathvariant="normal">4</mn><mo>+</mo></msubsup></mrow></math><span><svg:svg xmlns:svg="http://www.w3.org/2000/svg" width="24pt" height="15pt" class="svg-formula" dspmath="mathimg" md5hash="97c709e7ff43e05afce356dc3f53b497"><svg:image xmlns:xlink="http://www.w3.org/1999/xlink" xlink:href="acp-19-4419-2019-ie00001.svg" width="24pt" height="15pt" src="acp-19-4419-2019-ie00001.png"/></svg:svg></span></span>, as <span class="inline-formula">NH<sub>4</sub>Cl</span>, was measured at pH&thinsp;<span class="inline-formula">=</span>&thinsp;3 and found to be 1.2 (<span class="inline-formula">±0.1</span>)&thinsp;<span class="inline-formula">×</span>&thinsp;10<span class="inline-formula"><sup>−3</sup></span>&thinsp;M<span class="inline-formula"><sup>−1</sup></span>&thinsp;s<span class="inline-formula"><sup>−1</sup></span>, faster than the rate that would be estimated from rate measurements at much higher pHs. The solubilities of HNCO in the non-polar solvents <span class="inline-formula"><i>n</i></span>-octanol (<span class="inline-formula"><i>n</i></span>-<span class="inline-formula">C<sub>8</sub>H<sub>17</sub>OH</span>) and tridecane (<span class="inline-formula">C<sub>13</sub>H<sub>28</sub></span>) were found to be higher than aqueous solution for <span class="inline-formula"><i>n</i></span>-octanol (<span class="inline-formula">87±9</span>&thinsp;M&thinsp;atm<span class="inline-formula"><sup>−1</sup></span> at 298&thinsp;K) and much lower than aqueous solution for tridecane (<span class="inline-formula">1.7±0.17</span>&thinsp;M&thinsp;atm<span class="inline-formula"><sup>−1</sup></span> at 298&thinsp;K), features that have implications for multi-phase and membrane transport of HNCO. The first-order loss rate of HNCO in <span class="inline-formula"><i>n</i></span>-octanol was determined to be relatively slow, 5.7 (<span class="inline-formula">±1.4</span>)&thinsp;<span class="inline-formula">×</span>&thinsp;10<span class="inline-formula"><sup>−5</sup></span>&thinsp;s<span class="inline-formula"><sup>−1</sup></span>. The aqueous solubility of <span class="inline-formula">CH<sub>3</sub>NCO</span> was found to be 1.3 (<span class="inline-formula">±0.13</span>)&thinsp;M&thinsp;atm<span class="inline-formula"><sup>−1</sup></span> independent of pH, and <span class="inline-formula">CH<sub>3</sub>NCO</span> solubility in <span class="inline-formula"><i>n</i></span>-octanol was also determined at several temperatures and ranged from 4.0 (<span class="inline-formula">±0.5</span>)&thinsp;M&thinsp;atm<span class="inline-formula"><sup>−1</sup></span> at 298&thinsp;K to 2.8 (<span class="inline-formula">±0.3</span>)&thinsp;M&thinsp;atm<span class="inline-formula"><sup>−1</sup></span> at 310&thinsp;K. The aqueous hydrolysis of <span class="inline-formula">CH<sub>3</sub>NCO</span> was observed to be slightly acid-catalyzed, in agreement with literature values, and reactions with <span class="inline-formula"><i>n</i></span>-octanol ranged from 2.5 (<span class="inline-formula">±0.5</span>) to 5.3 (<span class="inline-formula">±0.7</span>)&thinsp;<span class="inline-formula">×</span>&thinsp;10<span class="inline-formula"><sup>−3</sup></span>&thinsp;s<span class="inline-formula"><sup>−1</sup></span> from 298 to 310&thinsp;K. The aqueous solubilities of XCN, determined at room temperature and neutral pH, were found to increase with halogen atom polarizability from 1.4 (<span class="inline-formula">±0.2</span>)&thinsp;M&thinsp;atm<span class="inline-formula"><sup>−1</sup></span> for ClCN and 8.2 (<span class="inline-formula">±0.8</span>)&thinsp;M&thinsp;atm<span class="inline-formula"><sup>−1</sup></span> for BrCN to 270 (<span class="inline-formula">±54</span>)&thinsp;M&thinsp;atm<span class="inline-formula"><sup>−1</sup></span> for ICN. Hydrolysis rates, where measurable, were in agreement with literature values. The atmospheric loss rates of HNCO, <span class="inline-formula">CH<sub>3</sub>NCO</span>, and XCN due to heterogeneous processes are estimated from solubilities and reaction rates. Lifetimes of HNCO range from about 1 day against deposition to neutral pH surfaces in the boundary layer, but otherwise can be as long as several months in the middle troposphere. The loss of <span class="inline-formula">CH<sub>3</sub>NCO</span> due to aqueous-phase processes is estimated to be slower than, or comparable to, the lifetime against OH reaction (3 months). The loss of XCNs due to aqueous uptake is estimated to range from being quite slow, with a lifetime of 2–6 months or more for ClCN and 1 week to 6 months for BrCN to 1 to 10 days for ICN. These characteristic times are shorter than photolysis lifetimes for ClCN and BrCN, implying that heterogeneous chemistry will be the controlling factor in their atmospheric removal. In contrast, the photolysis of ICN is estimated to be faster than heterogeneous loss for average midlatitude conditions.</p>
ISSN:1680-7316
1680-7324