Network


Latest external collaboration on country level. Dive into details by clicking on the dots.

Hotspot


Dive into the research topics where Jason M. St. Clair is active.

Publication


Featured researches published by Jason M. St. Clair.


Science | 2009

Unexpected Epoxide Formation in the Gas-Phase Photooxidation of Isoprene

Fabien Paulot; John D. Crounse; Henrik G. Kjaergaard; Andreas Kürten; Jason M. St. Clair; John H. Seinfeld; Paul O. Wennberg

No NO Isoprene, a five-carbon diene produced by plants, is the most abundant nonmethane hydrocarbon released into the atmosphere and plays an important role in tropospheric chemistry. Isoprene is also thought to affect climate by acting as a source of atmospheric aerosols. Paulot et al. (p. 730; see the Perspective by Kleindienst) now describe how isoprene may lead to the formation of secondary organic aerosols. In laboratory experiments, the photooxidation of isoprene in low-NO conditions, such as those which occur in vegetated regions far from anthropogenic influence, produced high yields of dihydroxy epoxides, a suspected precursor of the aerosols. This discovery could help to explain some of the more puzzling aspects of isoprene chemistry in remote regions. The oxidation of isoprenes in the absence of nitric oxide produces epoxides that can facilitate formation of organic aerosols. Emissions of nonmethane hydrocarbon compounds to the atmosphere from the biosphere exceed those from anthropogenic activity. Isoprene, a five-carbon diene, contributes more than 40% of these emissions. Once emitted to the atmosphere, isoprene is rapidly oxidized by the hydroxyl radical OH. We report here that under pristine conditions isoprene is oxidized primarily to hydroxyhydroperoxides. Further oxidation of these hydroxyhydroperoxides by OH leads efficiently to the formation of dihydroxyepoxides and OH reformation. Global simulations show an enormous flux—nearly 100 teragrams of carbon per year—of these epoxides to the atmosphere. The discovery of these highly soluble epoxides provides a missing link tying the gas-phase degradation of isoprene to the observed formation of organic aerosols.


Bulletin of the American Meteorological Society | 2015

The Deep Convective Clouds and Chemistry (DC3) Field Campaign

M. C. Barth; C. A. Cantrell; William H. Brune; Steven A. Rutledge; J. H. Crawford; Heidi Huntrieser; Lawrence D. Carey; Donald R. MacGorman; Morris L. Weisman; Kenneth E. Pickering; Eric C. Bruning; Bruce E. Anderson; Eric C. Apel; Michael I. Biggerstaff; Teresa L. Campos; Pedro Campuzano-Jost; R. C. Cohen; John D. Crounse; Douglas A. Day; Glenn S. Diskin; F. Flocke; Alan Fried; C. Garland; Brian G. Heikes; Shawn B. Honomichl; Rebecca S. Hornbrook; L. Gregory Huey; Jose L. Jimenez; Timothy J. Lang; Michael Lichtenstern

AbstractThe Deep Convective Clouds and Chemistry (DC3) field experiment produced an exceptional dataset on thunderstorms, including their dynamical, physical, and electrical structures and their impact on the chemical composition of the troposphere. The field experiment gathered detailed information on the chemical composition of the inflow and outflow regions of midlatitude thunderstorms in northeast Colorado, west Texas to central Oklahoma, and northern Alabama. A unique aspect of the DC3 strategy was to locate and sample the convective outflow a day after active convection in order to measure the chemical transformations within the upper-tropospheric convective plume. These data are being analyzed to investigate transport and dynamics of the storms, scavenging of soluble trace gases and aerosols, production of nitrogen oxides by lightning, relationships between lightning flash rates and storm parameters, chemistry in the upper troposphere that is affected by the convection, and related source character...


Atmospheric Chemistry and Physics | 2016

Why do models overestimate surface ozone in the Southeast United States

Katherine R. Travis; Daniel J. Jacob; Jenny A. Fisher; Patrick S. Kim; Eloise A. Marais; Lei Zhu; Karen Yu; Christopher Miller; Robert M. Yantosca; Melissa P. Sulprizio; Anne M. Thompson; Paul O. Wennberg; John D. Crounse; Jason M. St. Clair; R. C. Cohen; Joshua L. Laughner; Jack E. Dibb; Samuel R. Hall; Kirk Ullmann; G. M. Wolfe; I. B. Pollack; J. Peischl; J. A. Neuman; X. Zhou

Ozone pollution in the Southeast US involves complex chemistry driven by emissions of anthropogenic nitrogen oxide radicals (NOx ≡ NO + NO2) and biogenic isoprene. Model estimates of surface ozone concentrations tend to be biased high in the region and this is of concern for designing effective emission control strategies to meet air quality standards. We use detailed chemical observations from the SEAC4RS aircraft campaign in August and September 2013, interpreted with the GEOS-Chem chemical transport model at 0.25°×0.3125° horizontal resolution, to better understand the factors controlling surface ozone in the Southeast US. We find that the National Emission Inventory (NEI) for NOx from the US Environmental Protection Agency (EPA) is too high. This finding is based on SEAC4RS observations of NOx and its oxidation products, surface network observations of nitrate wet deposition fluxes, and OMI satellite observations of tropospheric NO2 columns. Our results indicate that NEI NOx emissions from mobile and industrial sources must be reduced by 30-60%, dependent on the assumption of the contribution by soil NOx emissions. Upper tropospheric NO2 from lightning makes a large contribution to satellite observations of tropospheric NO2 that must be accounted for when using these data to estimate surface NOx emissions. We find that only half of isoprene oxidation proceeds by the high-NOx pathway to produce ozone; this fraction is only moderately sensitive to changes in NOx emissions because isoprene and NOx emissions are spatially segregated. GEOS-Chem with reduced NOx emissions provides an unbiased simulation of ozone observations from the aircraft, and reproduces the observed ozone production efficiency in the boundary layer as derived from a regression of ozone and NOx oxidation products. However, the model is still biased high by 8±13 ppb relative to observed surface ozone in the Southeast US. Ozonesondes launched during midday hours show a 7 ppb ozone decrease from 1.5 km to the surface that GEOS-Chem does not capture. This bias may reflect a combination of excessive vertical mixing and net ozone production in the model boundary layer.


Proceedings of the National Academy of Sciences of the United States of America | 2015

Rapid Deposition of Oxidized Biogenic Compounds to a Temperate Forest

Tran B. Nguyen; John D. Crounse; Alex P. Teng; Jason M. St. Clair; Fabien Paulot; Glenn M. Wolfe; Paul O. Wennberg

Significance Dry deposition is an important removal mechanism for oxidized atmospheric compounds. This process remains, however, poorly understood due to the scarcity of direct flux observations for all but small, inorganic molecules in the atmosphere. The chemically speciated fluxes presented here comprise a unique and novel dataset that quantifies the dry deposition velocities for a variety of trace gases in a typical forested ecosystem. The data illustrate the key role of molecular diffusion in the atmosphere−biosphere exchange of water-soluble species. Furthermore, this work enabled evaluation of the dry deposition parameterization in a global chemical transport model. The results aid in resolving key discrepancies within the global model, resulting in more-accurate predictions of trace gas lifetimes and surface concentrations. We report fluxes and dry deposition velocities for 16 atmospheric compounds above a southeastern United States forest, including: hydrogen peroxide (H2O2), nitric acid (HNO3), hydrogen cyanide (HCN), hydroxymethyl hydroperoxide, peroxyacetic acid, organic hydroxy nitrates, and other multifunctional species derived from the oxidation of isoprene and monoterpenes. The data suggest that dry deposition is the dominant daytime sink for small, saturated oxygenates. Greater than 6 wt %C emitted as isoprene by the forest was returned by dry deposition of its oxidized products. Peroxides account for a large fraction of the oxidant flux, possibly eclipsing ozone in more pristine regions. The measured organic nitrates comprise a sizable portion (15%) of the oxidized nitrogen input into the canopy, with HNO3 making up the balance. We observe that water-soluble compounds (e.g., strong acids and hydroperoxides) deposit with low surface resistance whereas compounds with moderate solubility (e.g., organic nitrates and hydroxycarbonyls) or poor solubility (e.g., HCN) exhibited reduced uptake at the surface of plants. To first order, the relative deposition velocities of water-soluble compounds are constrained by their molecular diffusivity. From resistance modeling, we infer a substantial emission flux of formic acid at the canopy level (∼1 nmol m−2⋅s−1). GEOS−Chem, a widely used atmospheric chemical transport model, currently underestimates dry deposition for most molecules studied in this work. Reconciling GEOS−Chem deposition velocities with observations resulted in up to a 45% decrease in the simulated surface concentration of trace gases.


Journal of Physical Chemistry A | 2014

Gas Phase Production and Loss of Isoprene Epoxydiols

Kelvin H. Bates; John D. Crounse; Jason M. St. Clair; Nathan B. Bennett; Tran B. Nguyen; John H. Seinfeld; Brian M. Stoltz; Paul O. Wennberg

Isoprene epoxydiols (IEPOX) form in high yields from the OH-initiated oxidation of isoprene under low-NO conditions. These compounds contribute significantly to secondary organic aerosol formation. Their gas-phase chemistry has, however, remained largely unexplored. In this study, we characterize the formation of IEPOX isomers from the oxidation of isoprene by OH. We find that cis-β- and trans-β-IEPOX are the dominant isomers produced, and that they are created in an approximate ratio of 1:2 from the low-NO oxidation of isoprene. Three isomers of IEPOX, including cis-β- and trans-β, were synthesized and oxidized by OH in environmental chambers under high- and low-NO conditions. We find that IEPOX reacts with OH at 299 K with rate coefficients of (0.84 ± 0.07) × 10(-11), (1.52 ± 0.07) × 10(-11), and (0.98 ± 0.05) × 10(-11) cm(3) molecule(-1) s(-1) for the δ1, cis-β, and trans-β isomers. Finally, yields of the first-generation products of IEPOX + OH oxidation were measured, and a new mechanism of IEPOX oxidation is proposed here to account for the observed products. The substantial yield of glyoxal and methylglyoxal from IEPOX oxidation may help explain elevated levels of those compounds observed in low-NO environments with high isoprene emissions.


Environmental Science & Technology | 2015

Formation of Low Volatility Organic Compounds and Secondary Organic Aerosol from Isoprene Hydroxyhydroperoxide Low-NO Oxidation.

Jordan E. Krechmer; Matthew M. Coggon; Paola Massoli; Tran B. Nguyen; John D. Crounse; Weiwei Hu; Douglas A. Day; Geoffrey S. Tyndall; Daven K. Henze; Jean C. Rivera-Rios; J. B. Nowak; Joel R. Kimmel; Roy L. Mauldin; Harald Stark; John T. Jayne; Mikko Sipilä; Heikki Junninen; Jason M. St. Clair; Xuan Zhang; Philip A. Feiner; Li Zhang; David O. Miller; William H. Brune; Frank N. Keutsch; Paul O. Wennberg; John H. Seinfeld; Douglas R. Worsnop; Jose L. Jimenez; Manjula R. Canagaratna

Gas-phase low volatility organic compounds (LVOC), produced from oxidation of isoprene 4-hydroxy-3-hydroperoxide (4,3-ISOPOOH) under low-NO conditions, were observed during the FIXCIT chamber study. Decreases in LVOC directly correspond to appearance and growth in secondary organic aerosol (SOA) of consistent elemental composition, indicating that LVOC condense (at OA below 1 μg m(-3)). This represents the first simultaneous measurement of condensing low volatility species from isoprene oxidation in both the gas and particle phases. The SOA formation in this study is separate from previously described isoprene epoxydiol (IEPOX) uptake. Assigning all condensing LVOC signals to 4,3-ISOPOOH oxidation in the chamber study implies a wall-loss corrected non-IEPOX SOA mass yield of ∼4%. By contrast to monoterpene oxidation, in which extremely low volatility VOC (ELVOC) constitute the organic aerosol, in the isoprene system LVOC with saturation concentrations from 10(-2) to 10 μg m(-3) are the main constituents. These LVOC may be important for the growth of nanoparticles in environments with low OA concentrations. LVOC observed in the chamber were also observed in the atmosphere during SOAS-2013 in the Southeastern United States, with the expected diurnal cycle. This previously uncharacterized aerosol formation pathway could account for ∼5.0 Tg yr(-1) of SOA production, or 3.3% of global SOA.


Environmental Science & Technology | 2013

Observational insights into aerosol formation from isoprene.

David R. Worton; Jason D. Surratt; Brian W. Lafranchi; A. W. H. Chan; Yunliang Zhao; R. J. Weber; Jeong Hoo Park; J. B. Gilman; Joost A. de Gouw; Changhyoun Park; Gunnar W. Schade; Melinda R. Beaver; Jason M. St. Clair; John D. Crounse; Paul O. Wennberg; Glenn M. Wolfe; Sara Harrold; Joel A. Thornton; Delphine K. Farmer; Kenneth S. Docherty; Michael J. Cubison; Jose L. Jimenez; Amanda A. Frossard; Lynn M. Russell; Kasper Kristensen; Marianne Glasius; Jingqiu Mao; Xinrong Ren; William H. Brune; E. C. Browne

Atmospheric photooxidation of isoprene is an important source of secondary organic aerosol (SOA) and there is increasing evidence that anthropogenic oxidant emissions can enhance this SOA formation. In this work, we use ambient observations of organosulfates formed from isoprene epoxydiols (IEPOX) and methacrylic acid epoxide (MAE) and a broad suite of chemical measurements to investigate the relative importance of nitrogen oxide (NO/NO2) and hydroperoxyl (HO2) SOA formation pathways from isoprene at a forested site in California. In contrast to IEPOX, the calculated production rate of MAE was observed to be independent of temperature. This is the result of the very fast thermolysis of MPAN at high temperatures that affects the distribution of the MPAN reservoir (MPAN / MPA radical) reducing the fraction that can react with OH to form MAE and subsequently SOA (F(MAE formation)). The strong temperature dependence of F(MAE formation) helps to explain our observations of similar concentrations of IEPOX-derived organosulfates (IEPOX-OS; ~1 ng m(-3)) and MAE-derived organosulfates (MAE-OS; ~1 ng m(-3)) under cooler conditions (lower isoprene concentrations) and much higher IEPOX-OS (~20 ng m(-3)) relative to MAE-OS (<0.0005 ng m(-3)) at higher temperatures (higher isoprene concentrations). A kinetic model of IEPOX and MAE loss showed that MAE forms 10-100 times more ring-opening products than IEPOX and that both are strongly dependent on aerosol water content when aerosol pH is constant. However, the higher fraction of MAE ring opening products does not compensate for the lower MAE production under warmer conditions (higher isoprene concentrations) resulting in lower formation of MAE-derived products relative to IEPOX at the surface. In regions of high NOx, high isoprene emissions and strong vertical mixing the slower MPAN thermolysis rate aloft could increase the fraction of MPAN that forms MAE resulting in a vertically varying isoprene SOA source.


Review of Scientific Instruments | 2010

Chemical ionization tandem mass spectrometer for the in situ measurement of methyl hydrogen peroxide

Jason M. St. Clair; David C. McCabe; John D. Crounse; Urs Steiner; Paul O. Wennberg

A new approach for measuring gas-phase methyl hydrogen peroxide [(MHP) CH(3)OOH] utilizing chemical ionization mass spectrometry is presented. Tandem mass spectrometry is used to avoid mass interferences that hindered previous attempts to measure atmospheric CH(3)OOH with CF(3)O(-) clustering chemistry. CH(3)OOH has been successfully measured in situ using this technique during both airborne and ground-based campaigns. The accuracy and precision for the MHP measurement are a function of water vapor mixing ratio. Typical precision at 500 pptv MHP and 100 ppmv H(2)O is ±80 pptv (2 sigma) for a 1 s integration period. The accuracy at 100 ppmv H(2)O is estimated to be better than ±40%. Chemical ionization tandem mass spectrometry shows considerable promise for the determination of in situ atmospheric trace gas mixing ratios where isobaric compounds or mass interferences impede accurate measurements.


Journal of Physical Chemistry A | 2016

Kinetics and Products of the Reaction of the First-Generation Isoprene Hydroxy Hydroperoxide (ISOPOOH) with OH

Jason M. St. Clair; Jean C. Rivera-Rios; John D. Crounse; Hasse C. Knap; Kelvin H. Bates; Alex P. Teng; Solvejg Jørgensen; Henrik G. Kjaergaard; Frank N. Keutsch; Paul O. Wennberg

The atmospheric oxidation of isoprene by the OH radical leads to the formation of several isomers of an unsaturated hydroxy hydroperoxide, ISOPOOH. Oxidation of ISOPOOH by OH produces epoxydiols, IEPOX, which have been shown to contribute mass to secondary organic aerosol (SOA). We present kinetic rate constant measurements for OH + ISOPOOH using synthetic standards of the two major isomers: (1,2)- and (4,3)-ISOPOOH. At 297 K, the total OH rate constant is 7.5 ± 1.2 × 10(-11) cm(3) molecule(-1) s(-1) for (1,2)-ISOPOOH and 1.18 ± 0.19 × 10(-10) cm(3) molecule(-1) s(-1) for (4,3)-ISOPOOH. Abstraction of the hydroperoxy hydrogen accounts for approximately 12% and 4% of the reactivity for (1,2)-ISOPOOH and (4,3)-ISOPOOH, respectively. The sum of all H-abstractions account for approximately 15% and 7% of the reactivity for (1,2)-ISOPOOH and (4,3)-ISOPOOH, respectively. The major product observed from both ISOPOOH isomers was IEPOX (cis-β and trans-β isomers), with a ∼ 2:1 preference for trans-β IEPOX and similar total yields from each ISOPOOH isomer (∼ 70-80%). An IEPOX global production rate of more than 100 Tg C each year is estimated from this chemistry using a global 3D chemical transport model, similar to earlier estimates. Finally, following addition of OH to ISOPOOH, approximately 13% of the reactivity proceeds via addition of O2 at 297 K and 745 Torr. In the presence of NO, these peroxy radicals lead to formation of small carbonyl compounds. Under HO2 dominated chemistry, no products are observed from these channels. We suggest that the major products, highly oxygenated organic peroxides, are lost to the chamber walls. In the atmosphere, formation of these compounds may contribute to organic aerosol mass.


Atmospheric Chemistry and Physics | 2016

Organic nitrate chemistry and its implications for nitrogen budgets in an isoprene- and monoterpene-rich atmosphere: constraints from aircraft (SEAC 4 RS) and ground-based (SOAS) observations in the Southeast US

Jenny A. Fisher; Daniel J. Jacob; Katherine R. Travis; Patrick S. Kim; Eloise A. Marais; Christopher Miller; Karen Yu; Lei Zhu; Robert M. Yantosca; Melissa P. Sulprizio; Jingqiu Mao; Paul O. Wennberg; John D. Crounse; Alex P. Teng; Tran B. Nguyen; Jason M. St. Clair; R. C. Cohen; Paul M. Romer; Benjamin A. Nault; P. J. Wooldridge; Jose L. Jimenez; Pedro Campuzano-Jost; Douglas A. Day; Weiwei Hu; Paul B. Shepson; Fulizi Xiong; D. R. Blake; Allen H. Goldstein; Pawel K. Misztal; T. F. Hanisco

Formation of organic nitrates (RONO2) during oxidation of biogenic volatile organic compounds (BVOCs: isoprene, monoterpenes) is a significant loss pathway for atmospheric nitrogen oxide radicals (NOx), but the chemistry of RONO2 formation and degradation remains uncertain. Here we implement a new BVOC oxidation mechanism (including updated isoprene chemistry, new monoterpene chemistry, and particle uptake of RONO2) in the GEOS-Chem global chemical transport model with ∼25 × 25 km2 resolution over North America. We evaluate the model using aircraft (SEAC4RS) and ground-based (SOAS) observations of NOx, BVOCs, and RONO2 from the Southeast US in summer 2013. The updated simulation successfully reproduces the concentrations of individual gas- and particle-phase RONO2 species measured during the campaigns. Gas-phase isoprene nitrates account for 25-50% of observed RONO2 in surface air, and we find that another 10% is contributed by gas-phase monoterpene nitrates. Observations in the free troposphere show an important contribution from long-lived nitrates derived from anthropogenic VOCs. During both campaigns, at least 10% of observed boundary layer RONO2 were in the particle phase. We find that aerosol uptake followed by hydrolysis to HNO3 accounts for 60% of simulated gas-phase RONO2 loss in the boundary layer. Other losses are 20% by photolysis to recycle NOx and 15% by dry deposition. RONO2 production accounts for 20% of the net regional NOx sink in the Southeast US in summer, limited by the spatial segregation between BVOC and NOx emissions. This segregation implies that RONO2 production will remain a minor sink for NOx in the Southeast US in the future even as NOx emissions continue to decline.

Collaboration


Dive into the Jason M. St. Clair's collaboration.

Top Co-Authors

Avatar

John D. Crounse

California Institute of Technology

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

T. F. Hanisco

Goddard Space Flight Center

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Jose L. Jimenez

University of Colorado Boulder

View shared research outputs
Top Co-Authors

Avatar

D. R. Blake

University of California

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Researchain Logo
Decentralizing Knowledge