var bibbase_data = {"data":"\"Loading..\"\n\n
\n\n \n\n \n\n \n \n\n \n\n \n \n\n \n\n \n
\n generated by\n \n \"bibbase.org\"\n\n \n
\n \n\n
\n\n \n\n\n
\n\n Excellent! Next you can\n create a new website with this list, or\n embed it in an existing web page by copying & pasting\n any of the following snippets.\n\n
\n JavaScript\n (easiest)\n
\n \n <script src=\"https://bibbase.org/show?bib=https%3A%2F%2Fapi.zotero.org%2Fusers%2F4502912%2Fcollections%2F8BAJ5EJE%2Fitems%3Fkey%3DEM9YHr334UWQCR8ka10p48Y4%26format%3Dbibtex%26limit%3D100&jsonp=1&jsonp=1\"></script>\n \n
\n\n PHP\n
\n \n <?php\n $contents = file_get_contents(\"https://bibbase.org/show?bib=https%3A%2F%2Fapi.zotero.org%2Fusers%2F4502912%2Fcollections%2F8BAJ5EJE%2Fitems%3Fkey%3DEM9YHr334UWQCR8ka10p48Y4%26format%3Dbibtex%26limit%3D100&jsonp=1\");\n print_r($contents);\n ?>\n \n
\n\n iFrame\n (not recommended)\n
\n \n <iframe src=\"https://bibbase.org/show?bib=https%3A%2F%2Fapi.zotero.org%2Fusers%2F4502912%2Fcollections%2F8BAJ5EJE%2Fitems%3Fkey%3DEM9YHr334UWQCR8ka10p48Y4%26format%3Dbibtex%26limit%3D100&jsonp=1\"></iframe>\n \n
\n\n

\n For more details see the documention.\n

\n
\n
\n\n
\n\n This is a preview! To use this list on your own web site\n or create a new web site from it,\n create a free account. The file will be added\n and you will be able to edit it in the File Manager.\n We will show you instructions once you've created your account.\n
\n\n
\n\n

To the site owner:

\n\n

Action required! Mendeley is changing its\n API. In order to keep using Mendeley with BibBase past April\n 14th, you need to:\n

    \n
  1. renew the authorization for BibBase on Mendeley, and
  2. \n
  3. update the BibBase URL\n in your page the same way you did when you initially set up\n this page.\n
  4. \n
\n

\n\n

\n \n \n Fix it now\n

\n
\n\n
\n\n\n
\n \n \n
\n
\n  \n 2023\n \n \n (3)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Sulfuric Acid Nucleation Potential Model Applied to Complex Reacting Systems in the Atmosphere.\n \n \n \n \n\n\n \n Johnson, J. S.; and Jen, C. N.\n\n\n \n\n\n\n Journal of Geophysical Research: Atmospheres, 128(20): e2023JD039344. 2023.\n _eprint: https://agupubs.onlinelibrary.wiley.com/doi/pdf/10.1029/2023JD039344\n\n\n\n
\n\n\n\n \n \n \"SulfuricPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n  \n \n 2 downloads\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{johnson_sulfuric_2023,\n\ttitle = {Sulfuric {Acid} {Nucleation} {Potential} {Model} {Applied} to {Complex} {Reacting} {Systems} in the {Atmosphere}},\n\tvolume = {128},\n\tcopyright = {© 2023. The Authors.},\n\tissn = {2169-8996},\n\turl = {https://onlinelibrary.wiley.com/doi/abs/10.1029/2023JD039344},\n\tdoi = {10.1029/2023JD039344},\n\tabstract = {Atmospheric aerosol particles impact Earth's radiation balance by acting as seeds for cloud droplet formation. Over half of global cloud seed particles are formed by nucleation, a process where gas-phase compounds react to form stable particles. Reactions of sulfuric acid (SA) with a wide variety of atmospheric compounds have been previously shown to drive nucleation in the lower troposphere. However, global climate models poorly predict particle nucleation rates since current nucleation models do not describe nucleation for systems containing tens to hundreds of precursor compounds. The nucleation potential model (NPM) was recently developed to model SA nucleation of complex mixtures by measuring an effective base concentration using a 1-nm condensation particle counter. This technique for estimating particle nucleation rates can be deployed at a much higher spatial and temporal resolution than current methods which require detailed knowledge of all nucleation reactions and measurements, typically using a mass spectrometer, of all nucleation precursor gases. This work expands NPM by showing that this model can capture enhancement and suppression of SA nucleation rates within a complex mixture of organic and inorganic acids, ambient air, and across a range of atmospherically relevant relative humidities. In addition, an expression for calculating atmospheric nucleation rates was also derived from the NPM. Ultimately, NPM provides a simple way to measure and model the extent compounds in a complex mixture enhance SA nucleation rates using a condensation particle counter.},\n\tlanguage = {en},\n\tnumber = {20},\n\turldate = {2023-10-15},\n\tjournal = {Journal of Geophysical Research: Atmospheres},\n\tauthor = {Johnson, J. S. and Jen, C. N.},\n\tyear = {2023},\n\tnote = {\\_eprint: https://agupubs.onlinelibrary.wiley.com/doi/pdf/10.1029/2023JD039344},\n\tkeywords = {amines, methanesulfonic acid, nucleation, nucleation potential model, organic acids, sulfuric acid},\n\tpages = {e2023JD039344},\n}\n\n
\n
\n\n\n
\n Atmospheric aerosol particles impact Earth's radiation balance by acting as seeds for cloud droplet formation. Over half of global cloud seed particles are formed by nucleation, a process where gas-phase compounds react to form stable particles. Reactions of sulfuric acid (SA) with a wide variety of atmospheric compounds have been previously shown to drive nucleation in the lower troposphere. However, global climate models poorly predict particle nucleation rates since current nucleation models do not describe nucleation for systems containing tens to hundreds of precursor compounds. The nucleation potential model (NPM) was recently developed to model SA nucleation of complex mixtures by measuring an effective base concentration using a 1-nm condensation particle counter. This technique for estimating particle nucleation rates can be deployed at a much higher spatial and temporal resolution than current methods which require detailed knowledge of all nucleation reactions and measurements, typically using a mass spectrometer, of all nucleation precursor gases. This work expands NPM by showing that this model can capture enhancement and suppression of SA nucleation rates within a complex mixture of organic and inorganic acids, ambient air, and across a range of atmospherically relevant relative humidities. In addition, an expression for calculating atmospheric nucleation rates was also derived from the NPM. Ultimately, NPM provides a simple way to measure and model the extent compounds in a complex mixture enhance SA nucleation rates using a condensation particle counter.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Limited Role of Malonic Acid in Sulfuric Acid–Dimethylamine New Particle Formation.\n \n \n \n \n\n\n \n Fomete, S. K.; Kubečka, J.; Elm, J.; and Jen, C. N.\n\n\n \n\n\n\n ACS Omega, 8(22): 19807–19815. June 2023.\n Publisher: American Chemical Society\n\n\n\n
\n\n\n\n \n \n \"LimitedPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n  \n \n 2 downloads\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fomete_limited_2023,\n\ttitle = {Limited {Role} of {Malonic} {Acid} in {Sulfuric} {Acid}–{Dimethylamine} {New} {Particle} {Formation}},\n\tvolume = {8},\n\turl = {https://doi.org/10.1021/acsomega.3c01643},\n\tdoi = {10.1021/acsomega.3c01643},\n\tabstract = {Aerosols play an important role in climate and air quality; however, the mechanisms behind aerosol particle formation in the atmosphere are poorly understood. Studies have identified sulfuric acid, water, oxidized organics, and ammonia/amines as key precursors for forming aerosol particles in the atmosphere. Theoretical and experimental investigations have indicated that other species, such as organic acids, may be involved in atmospheric nucleation and growth of freshly formed aerosol particles. Organic acids, such as dicarboxylic acids, which are abundant in the atmosphere, have been measured in ultrafine aerosol particles. These observations suggest that organic acids may contribute to new particle formation in the atmosphere but their role remains ambiguous. This study examines how malonic acid interacts with sulfuric acid and dimethylamine to form new particles at warm boundary layer conditions using experimental observations from a laminar flow reactor and quantum chemical calculations coupled with cluster dynamics simulations. Observations reveal that malonic acid does not contribute to the initial steps (formation of {\\textless}1 nm diameter particle) of nucleation with sulfuric acid-dimethylamine. In addition, malonic acid was found to not participate in the subsequent growth of the freshly nucleated 1 nm particles from sulfuric acid-dimethylamine reactions to diameters of 2 nm.},\n\tnumber = {22},\n\turldate = {2023-08-16},\n\tjournal = {ACS Omega},\n\tauthor = {Fomete, Sandra K.W. and Kubečka, Jakub and Elm, Jonas and Jen, Coty N.},\n\tmonth = jun,\n\tyear = {2023},\n\tnote = {Publisher: American Chemical Society},\n\tpages = {19807--19815},\n}\n\n
\n
\n\n\n
\n Aerosols play an important role in climate and air quality; however, the mechanisms behind aerosol particle formation in the atmosphere are poorly understood. Studies have identified sulfuric acid, water, oxidized organics, and ammonia/amines as key precursors for forming aerosol particles in the atmosphere. Theoretical and experimental investigations have indicated that other species, such as organic acids, may be involved in atmospheric nucleation and growth of freshly formed aerosol particles. Organic acids, such as dicarboxylic acids, which are abundant in the atmosphere, have been measured in ultrafine aerosol particles. These observations suggest that organic acids may contribute to new particle formation in the atmosphere but their role remains ambiguous. This study examines how malonic acid interacts with sulfuric acid and dimethylamine to form new particles at warm boundary layer conditions using experimental observations from a laminar flow reactor and quantum chemical calculations coupled with cluster dynamics simulations. Observations reveal that malonic acid does not contribute to the initial steps (formation of \\textless1 nm diameter particle) of nucleation with sulfuric acid-dimethylamine. In addition, malonic acid was found to not participate in the subsequent growth of the freshly nucleated 1 nm particles from sulfuric acid-dimethylamine reactions to diameters of 2 nm.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Role of Methanesulfonic Acid in Sulfuric Acid–Amine and Ammonia New Particle Formation.\n \n \n \n \n\n\n \n Johnson, J. S.; and Jen, C. N.\n\n\n \n\n\n\n ACS Earth and Space Chemistry. March 2023.\n Publisher: American Chemical Society\n\n\n\n
\n\n\n\n \n \n \"RolePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n  \n \n 2 downloads\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{johnson_role_2023,\n\ttitle = {Role of {Methanesulfonic} {Acid} in {Sulfuric} {Acid}–{Amine} and {Ammonia} {New} {Particle} {Formation}},\n\turl = {https://doi.org/10.1021/acsearthspacechem.3c00017},\n\tdoi = {10.1021/acsearthspacechem.3c00017},\n\tabstract = {Aerosol nucleation accounts for over half of all seed particles for cloud droplet formation. In the atmosphere, sulfuric acid (SA) nucleates with ammonia, amines, oxidized organics, and many more compounds to form particles. Studies have also shown that methanesulfonic acid (MSA) nucleates independently with amines and ammonia. MSA and SA are produced simultaneously via dimethyl sulfide oxidation in the marine atmosphere. However, limited knowledge exists on how MSA and SA nucleate together in the presence of various atmospherically relevant base compounds, which is critical to predicting marine nucleation rates accurately. This work provides experimental evidence that SA and MSA react to form particles with amines and that the SA-MSA-base nucleation has different reaction pathways than SA-base nucleation. Specifically, the formation of the SA-MSA heterodimer creates more energetically favorable pathways for SA-MSA-methylamine nucleation and an enhancement of nucleation rates. However, SA-trimethylamine nucleation is suppressed by MSA, likely due to the steric hindrance of the MSA and trimethylamine. These results display the importance of including nucleation reactions between SA, MSA, and various amines to predict particle nucleation rates in the marine atmosphere.},\n\turldate = {2023-03-09},\n\tjournal = {ACS Earth and Space Chemistry},\n\tauthor = {Johnson, Jack S. and Jen, Coty N.},\n\tmonth = mar,\n\tyear = {2023},\n\tnote = {Publisher: American Chemical Society},\n}\n\n
\n
\n\n\n
\n Aerosol nucleation accounts for over half of all seed particles for cloud droplet formation. In the atmosphere, sulfuric acid (SA) nucleates with ammonia, amines, oxidized organics, and many more compounds to form particles. Studies have also shown that methanesulfonic acid (MSA) nucleates independently with amines and ammonia. MSA and SA are produced simultaneously via dimethyl sulfide oxidation in the marine atmosphere. However, limited knowledge exists on how MSA and SA nucleate together in the presence of various atmospherically relevant base compounds, which is critical to predicting marine nucleation rates accurately. This work provides experimental evidence that SA and MSA react to form particles with amines and that the SA-MSA-base nucleation has different reaction pathways than SA-base nucleation. Specifically, the formation of the SA-MSA heterodimer creates more energetically favorable pathways for SA-MSA-methylamine nucleation and an enhancement of nucleation rates. However, SA-trimethylamine nucleation is suppressed by MSA, likely due to the steric hindrance of the MSA and trimethylamine. These results display the importance of including nucleation reactions between SA, MSA, and various amines to predict particle nucleation rates in the marine atmosphere.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2022\n \n \n (6)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Ion–Molecule Rate Constants for Reactions of Sulfuric Acid with Acetate and Nitrate Ions.\n \n \n \n \n\n\n \n Fomete, S. K. W.; Johnson, J. S.; Myllys, N.; Neefjes, I.; Reischl, B.; and Jen, C. N.\n\n\n \n\n\n\n The Journal of Physical Chemistry A. October 2022.\n Publisher: American Chemical Society\n\n\n\n
\n\n\n\n \n \n \"Ion–MoleculePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n  \n \n 1 download\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fomete_ionmolecule_2022,\n\ttitle = {Ion–{Molecule} {Rate} {Constants} for {Reactions} of {Sulfuric} {Acid} with {Acetate} and {Nitrate} {Ions}},\n\tcopyright = {All rights reserved},\n\tissn = {1089-5639},\n\turl = {https://doi.org/10.1021/acs.jpca.2c02072},\n\tdoi = {10.1021/acs.jpca.2c02072},\n\tabstract = {Atmospheric nucleation from precursor gases is a significant source of cloud condensation nuclei in the troposphere and thus can affect the Earth’s radiative balance. Sulfuric acid, ammonia, and amines have been identified as key nucleation precursors in the atmosphere. Studies have also shown that atmospheric ions can react with sulfuric acid to form stable clusters in a process referred to as ion-induced nucleation (IIN). IIN follows similar reaction pathways as chemical ionization, which is used to detect and measure nucleation precursors via atmospheric pressure chemical ionization mass spectrometers. The rate at which ions form clusters depends on the ion–molecule rate constant. However, the rate constant varies based on the ion composition, which is often not known in the atmosphere. Previous studies have examined ion–molecule rate constants for sulfuric acid and nitrate ions but not for other atmospherically relevant ions like acetate. We report the relative rate constants of ion–molecule reactions between nitrate and acetate ions reacting with sulfuric acid. The ion–molecule rate constant for acetate and sulfuric acid is estimated to be a factor of 1.9–2.4 times higher than that of the known rate constant for nitrate and sulfuric acid. Using quantum chemistry, we find that acetate has a higher dipole moment and polarizability than nitrate. This may contribute to an increase in the collision cross-sectional area between acetate and sulfuric acid and lead to a greater reaction rate constant than nitrate. The ion–molecule rate constant for acetate with sulfuric acid will help quantify the contribution of acetate ions to atmospheric ion-induced new particle formation.},\n\turldate = {2022-10-31},\n\tjournal = {The Journal of Physical Chemistry A},\n\tauthor = {Fomete, Sandra K. W. and Johnson, Jack S. and Myllys, Nanna and Neefjes, Ivo and Reischl, Bernhard and Jen, Coty N.},\n\tmonth = oct,\n\tyear = {2022},\n\tnote = {Publisher: American Chemical Society},\n}\n\n
\n
\n\n\n
\n Atmospheric nucleation from precursor gases is a significant source of cloud condensation nuclei in the troposphere and thus can affect the Earth’s radiative balance. Sulfuric acid, ammonia, and amines have been identified as key nucleation precursors in the atmosphere. Studies have also shown that atmospheric ions can react with sulfuric acid to form stable clusters in a process referred to as ion-induced nucleation (IIN). IIN follows similar reaction pathways as chemical ionization, which is used to detect and measure nucleation precursors via atmospheric pressure chemical ionization mass spectrometers. The rate at which ions form clusters depends on the ion–molecule rate constant. However, the rate constant varies based on the ion composition, which is often not known in the atmosphere. Previous studies have examined ion–molecule rate constants for sulfuric acid and nitrate ions but not for other atmospherically relevant ions like acetate. We report the relative rate constants of ion–molecule reactions between nitrate and acetate ions reacting with sulfuric acid. The ion–molecule rate constant for acetate and sulfuric acid is estimated to be a factor of 1.9–2.4 times higher than that of the known rate constant for nitrate and sulfuric acid. Using quantum chemistry, we find that acetate has a higher dipole moment and polarizability than nitrate. This may contribute to an increase in the collision cross-sectional area between acetate and sulfuric acid and lead to a greater reaction rate constant than nitrate. The ion–molecule rate constant for acetate with sulfuric acid will help quantify the contribution of acetate ions to atmospheric ion-induced new particle formation.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Dilution and photooxidation driven processes explain the evolution of organic aerosol in wildfire plumes.\n \n \n \n \n\n\n \n Akherati, A.; He, Y.; Garofalo, L. A.; Hodshire, A. L.; Farmer, D. K.; Kreidenweis, S. M.; Permar, W.; Hu, L.; Fischer, E. V.; Jen, C. N.; Goldstein, A. H.; Levin, E. J. T.; DeMott, P. J.; Campos, T. L.; Flocke, F.; Reeves, J. M.; Toohey, D. W.; Pierce, J. R.; and Jathar, S. H.\n\n\n \n\n\n\n Environmental Science: Atmospheres, 2(5): 1000–1022. September 2022.\n Publisher: RSC\n\n\n\n
\n\n\n\n \n \n \"DilutionPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n  \n \n 1 download\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{akherati_dilution_2022,\n\ttitle = {Dilution and photooxidation driven processes explain the evolution of organic aerosol in wildfire plumes},\n\tvolume = {2},\n\tissn = {2634-3606},\n\turl = {http://pubs.rsc.org/en/content/articlelanding/2022/ea/d1ea00082a},\n\tdoi = {10.1039/D1EA00082A},\n\tabstract = {Wildfires are an important atmospheric source of primary organic aerosol (POA) and precursors for secondary organic aerosol (SOA) at regional and global scales. However, there are large uncertainties surrounding the emissions and physicochemical processes that control the transformation, evolution, and properties of POA and SOA in large wildfire plumes. We develop a plume version of a kinetic model to simulate the dilution, oxidation chemistry, thermodynamic properties, and microphysics of organic aerosol (OA) in wildfire smoke. The model is applied to study the in-plume OA in four large wildfire smoke plumes intercepted during an aircraft-based field campaign in summer 2018 in the western United States. Based on estimates of dilution and oxidant concentrations before the aircraft first intercepted the plumes, we simulate the OA evolution from very close to the fire to several hours downwind. Our model results and sensitivity simulations suggest that dilution-driven evaporation of POA and simultaneous photochemical production of SOA are likely to explain the observed evolution in OA mass with physical age. The model, however, substantially underestimates the change in the oxygen-to-carbon ratio of the OA compared to measurements. In addition, we show that the rapid chemical transformation within the first hour after emission is driven by higher-than-ambient OH concentrations (3 × 106–107 molecules per cm3) and the slower evolution over the next several hours is a result of lower-than-ambient OH concentrations ({\\textless}106 molecules per cm3) and depleted SOA precursors. Model predictions indicate that the OA measured several hours downwind of the fire is still dominated by POA but with an SOA fraction that varies between 30\\% and 56\\% of the total OA. Semivolatile, heterocyclic, and oxygenated aromatic compounds, in that order, were found to contribute substantially ({\\textgreater}90\\%) to SOA formation. Future work needs to focus on better understanding the dynamic evolution closer to the fire and resolving the rapid change in the oxidation state of OA with physical age.},\n\tlanguage = {en},\n\tnumber = {5},\n\turldate = {2022-10-24},\n\tjournal = {Environmental Science: Atmospheres},\n\tauthor = {Akherati, Ali and He, Yicong and Garofalo, Lauren A. and Hodshire, Anna L. and Farmer, Delphine K. and Kreidenweis, Sonia M. and Permar, Wade and Hu, Lu and Fischer, Emily V. and Jen, Coty N. and Goldstein, Allen H. and Levin, Ezra J. T. and DeMott, Paul J. and Campos, Teresa L. and Flocke, Frank and Reeves, John M. and Toohey, Darin W. and Pierce, Jeffrey R. and Jathar, Shantanu H.},\n\tmonth = sep,\n\tyear = {2022},\n\tnote = {Publisher: RSC},\n\tpages = {1000--1022},\n}\n\n
\n
\n\n\n
\n Wildfires are an important atmospheric source of primary organic aerosol (POA) and precursors for secondary organic aerosol (SOA) at regional and global scales. However, there are large uncertainties surrounding the emissions and physicochemical processes that control the transformation, evolution, and properties of POA and SOA in large wildfire plumes. We develop a plume version of a kinetic model to simulate the dilution, oxidation chemistry, thermodynamic properties, and microphysics of organic aerosol (OA) in wildfire smoke. The model is applied to study the in-plume OA in four large wildfire smoke plumes intercepted during an aircraft-based field campaign in summer 2018 in the western United States. Based on estimates of dilution and oxidant concentrations before the aircraft first intercepted the plumes, we simulate the OA evolution from very close to the fire to several hours downwind. Our model results and sensitivity simulations suggest that dilution-driven evaporation of POA and simultaneous photochemical production of SOA are likely to explain the observed evolution in OA mass with physical age. The model, however, substantially underestimates the change in the oxygen-to-carbon ratio of the OA compared to measurements. In addition, we show that the rapid chemical transformation within the first hour after emission is driven by higher-than-ambient OH concentrations (3 × 106–107 molecules per cm3) and the slower evolution over the next several hours is a result of lower-than-ambient OH concentrations (\\textless106 molecules per cm3) and depleted SOA precursors. Model predictions indicate that the OA measured several hours downwind of the fire is still dominated by POA but with an SOA fraction that varies between 30% and 56% of the total OA. Semivolatile, heterocyclic, and oxygenated aromatic compounds, in that order, were found to contribute substantially (\\textgreater90%) to SOA formation. Future work needs to focus on better understanding the dynamic evolution closer to the fire and resolving the rapid change in the oxidation state of OA with physical age.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Preventing spread of aerosolized infectious particles during medical procedures: A lab-based analysis of an inexpensive plastic enclosure.\n \n \n \n \n\n\n \n Monroe, L. W.; Johnson, J. S.; Gutstein, H. B.; Lawrence, J. P.; Lejeune, K.; Sullivan, R. C.; and Jen, C. N.\n\n\n \n\n\n\n PLOS ONE, 17(9): e0273194. September 2022.\n Publisher: Public Library of Science\n\n\n\n
\n\n\n\n \n \n \"PreventingPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n  \n \n 2 downloads\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{monroe_preventing_2022,\n\ttitle = {Preventing spread of aerosolized infectious particles during medical procedures: {A} lab-based analysis of an inexpensive plastic enclosure},\n\tvolume = {17},\n\tissn = {1932-6203},\n\tshorttitle = {Preventing spread of aerosolized infectious particles during medical procedures},\n\turl = {https://journals.plos.org/plosone/article?id=10.1371/journal.pone.0273194},\n\tdoi = {10.1371/journal.pone.0273194},\n\tabstract = {Severe viral respiratory diseases, such as SARS-CoV-2, are transmitted through aerosol particles produced by coughing, talking, and breathing. Medical procedures including tracheal intubation, extubation, dental work, and any procedure involving close contact with a patient’s airways can increase exposure to infectious aerosol particles. This presents a significant risk for viral exposure of nearby healthcare workers during and following patient care. Previous studies have examined the effectiveness of plastic enclosures for trapping aerosol particles and protecting health-care workers. However, many of these enclosures are expensive or are burdensome for healthcare workers to work with. In this study, a low-cost plastic enclosure was designed to reduce aerosol spread and viral transmission during medical procedures, while also alleviating issues found in the design and use of other medical enclosures to contain aerosols. This enclosure is fabricated from clear polycarbonate for maximum visibility. A large single-side cutout provides health care providers with ease of access to the patient with a separate cutout for equipment access. A survey of medical providers in a local hospital network demonstrated their approval of the enclosure’s ease of use and design. The enclosure with appropriate plastic covers reduced total escaped particle number concentrations (diameter {\\textgreater} 0.01 μm) by over 93\\% at 8 cm away from all openings. Concentration decay experiments indicated that the enclosure without active suction should be left on the patient for 15–20 minutes following a tracheal manipulation to allow sufficient time for {\\textgreater}90\\% of aerosol particles to settle upon interior surfaces. This decreases to 5 minutes when 30 LPM suction is applied. This enclosure is an inexpensive, easily implemented additional layer of protection that can be used to help contain infectious or otherwise potentially hazardous aerosol particles while providing access into the enclosure.},\n\tlanguage = {en},\n\tnumber = {9},\n\turldate = {2022-09-26},\n\tjournal = {PLOS ONE},\n\tauthor = {Monroe, Luke W. and Johnson, Jack S. and Gutstein, Howard B. and Lawrence, John P. and Lejeune, Keith and Sullivan, Ryan C. and Jen, Coty N.},\n\tmonth = sep,\n\tyear = {2022},\n\tnote = {Publisher: Public Library of Science},\n\tkeywords = {Aerosols, Coughing, Hands, Health care providers, Intubation, Medical devices and equipment, Medical risk factors, Safety equipment},\n\tpages = {e0273194},\n}\n\n
\n
\n\n\n
\n Severe viral respiratory diseases, such as SARS-CoV-2, are transmitted through aerosol particles produced by coughing, talking, and breathing. Medical procedures including tracheal intubation, extubation, dental work, and any procedure involving close contact with a patient’s airways can increase exposure to infectious aerosol particles. This presents a significant risk for viral exposure of nearby healthcare workers during and following patient care. Previous studies have examined the effectiveness of plastic enclosures for trapping aerosol particles and protecting health-care workers. However, many of these enclosures are expensive or are burdensome for healthcare workers to work with. In this study, a low-cost plastic enclosure was designed to reduce aerosol spread and viral transmission during medical procedures, while also alleviating issues found in the design and use of other medical enclosures to contain aerosols. This enclosure is fabricated from clear polycarbonate for maximum visibility. A large single-side cutout provides health care providers with ease of access to the patient with a separate cutout for equipment access. A survey of medical providers in a local hospital network demonstrated their approval of the enclosure’s ease of use and design. The enclosure with appropriate plastic covers reduced total escaped particle number concentrations (diameter \\textgreater 0.01 μm) by over 93% at 8 cm away from all openings. Concentration decay experiments indicated that the enclosure without active suction should be left on the patient for 15–20 minutes following a tracheal manipulation to allow sufficient time for \\textgreater90% of aerosol particles to settle upon interior surfaces. This decreases to 5 minutes when 30 LPM suction is applied. This enclosure is an inexpensive, easily implemented additional layer of protection that can be used to help contain infectious or otherwise potentially hazardous aerosol particles while providing access into the enclosure.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Experimental and Theoretical Study on the Enhancement of Alkanolamines on Sulfuric Acid Nucleation.\n \n \n \n \n\n\n \n Fomete, S. K. W.; Johnson, J. S.; Myllys, N.; and Jen, C. N.\n\n\n \n\n\n\n The Journal of Physical Chemistry A, 126(25): 4057–4067. June 2022.\n Publisher: American Chemical Society\n\n\n\n
\n\n\n\n \n \n \"ExperimentalPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n  \n \n 1 download\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fomete_experimental_2022,\n\ttitle = {Experimental and {Theoretical} {Study} on the {Enhancement} of {Alkanolamines} on {Sulfuric} {Acid} {Nucleation}},\n\tvolume = {126},\n\tcopyright = {All rights reserved},\n\tissn = {1089-5639},\n\turl = {https://doi.org/10.1021/acs.jpca.2c01672},\n\tdoi = {10.1021/acs.jpca.2c01672},\n\tabstract = {Alkanolamines such as monoethanolamine (MEA), diethanolamine (DEA), and triethanolamine (TEA) are extensively used for CO2 capture and consumer products. Despite their prevalence in industrial applications, the fate of alkanolamines in the atmosphere remains relatively unknown. One likely reaction pathway for these chemicals in the atmosphere is new particle formation with sulfuric acid. Here, we present the first experimental results showing the formation of sulfuric acid dimers enhanced by MEA, DEA, and TEA from the measurement of molecular clusters. This study examines the nucleation reactions of MEA, DEA, and TEA with sulfuric acid in a clean, laminar flow reactor. The chemical compositions and concentrations of the freshly nucleated clusters were analyzed using a custom-built atmospheric pressure chemical ionization long time-of-flight mass spectrometer known as the Pittsburgh Cluster CIMS. Quantum chemical calculations and kinetic modeling of sulfuric acid-MEA/DEA/TEA clusters were also performed to determine relative cluster stabilities between these sulfuric acid–base systems. Experimental results indicate that MEA, DEA, and TEA at the part per trillion by volume (pptv) concentrations can enhance sulfuric acid dimer formation rates but to a lesser extent than dimethylamine (DMA). Thus, MEA, DEA, and TEA will potentially play an important role in new particle formation in industrial cities where these alkanolamines are emitted.},\n\tnumber = {25},\n\turldate = {2022-06-30},\n\tjournal = {The Journal of Physical Chemistry A},\n\tauthor = {Fomete, Sandra K. W. and Johnson, Jack S. and Myllys, Nanna and Jen, Coty N.},\n\tmonth = jun,\n\tyear = {2022},\n\tnote = {Publisher: American Chemical Society},\n\tpages = {4057--4067},\n}\n\n
\n
\n\n\n
\n Alkanolamines such as monoethanolamine (MEA), diethanolamine (DEA), and triethanolamine (TEA) are extensively used for CO2 capture and consumer products. Despite their prevalence in industrial applications, the fate of alkanolamines in the atmosphere remains relatively unknown. One likely reaction pathway for these chemicals in the atmosphere is new particle formation with sulfuric acid. Here, we present the first experimental results showing the formation of sulfuric acid dimers enhanced by MEA, DEA, and TEA from the measurement of molecular clusters. This study examines the nucleation reactions of MEA, DEA, and TEA with sulfuric acid in a clean, laminar flow reactor. The chemical compositions and concentrations of the freshly nucleated clusters were analyzed using a custom-built atmospheric pressure chemical ionization long time-of-flight mass spectrometer known as the Pittsburgh Cluster CIMS. Quantum chemical calculations and kinetic modeling of sulfuric acid-MEA/DEA/TEA clusters were also performed to determine relative cluster stabilities between these sulfuric acid–base systems. Experimental results indicate that MEA, DEA, and TEA at the part per trillion by volume (pptv) concentrations can enhance sulfuric acid dimer formation rates but to a lesser extent than dimethylamine (DMA). Thus, MEA, DEA, and TEA will potentially play an important role in new particle formation in industrial cities where these alkanolamines are emitted.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n A sulfuric acid nucleation potential model for the atmosphere.\n \n \n \n \n\n\n \n Johnson, J. S.; and Jen, C. N.\n\n\n \n\n\n\n Atmospheric Chemistry and Physics, 22(12): 8287–8297. June 2022.\n Publisher: Copernicus GmbH\n\n\n\n
\n\n\n\n \n \n \"APaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n  \n \n 1 download\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{johnson_sulfuric_2022,\n\ttitle = {A sulfuric acid nucleation potential model for the atmosphere},\n\tvolume = {22},\n\tcopyright = {All rights reserved},\n\tissn = {1680-7316},\n\turl = {https://acp.copernicus.org/articles/22/8287/2022/},\n\tdoi = {10.5194/acp-22-8287-2022},\n\tabstract = {{\\textless}p{\\textgreater}{\\textless}strong class="journal-contentHeaderColor"{\\textgreater}Abstract.{\\textless}/strong{\\textgreater} Observations over the last decade have demonstrated that the atmosphere contains potentially hundreds of compounds that can react with sulfuric acid to nucleate stable aerosol particles. Consequently, modeling atmospheric nucleation requires detailed knowledge of nucleation reaction kinetics and spatially and temporally resolved measurements of numerous precursor compounds. This study introduces the Nucleation Potential Model (NPM), a novel nucleation model that dramatically simplifies the diverse reactions between sulfuric acid and any combination of precursor gases. The NPM predicts 1 nm nucleation rates from only two measurable gas concentrations, regardless of whether all precursor gases are known. The NPM describes sulfuric acid nucleating with a parameterized base compound at an effective base concentration, [{\\textless}span class="inline-formula"{\\textgreater}\\textit{B}$_{\\textrm{eff}}${\\textless}/span{\\textgreater}]. [{\\textless}span class="inline-formula"{\\textgreater}\\textit{B}$_{\\textrm{eff}}${\\textless}/span{\\textgreater}] captures the ability of a compound or mixture to form stable clusters with sulfuric acid and is estimated from measured 1 nm particle concentrations. The NPM is applied to experimental and field observations of sulfuric acid nucleation to demonstrate how [{\\textless}span class="inline-formula"{\\textgreater}\\textit{B}$_{\\textrm{eff}}${\\textless}/span{\\textgreater}] varies for different stabilizing compounds, mixtures, and sampling locations. Analysis of previous field observations shows distinct differences in [{\\textless}span class="inline-formula"{\\textgreater}\\textit{B}$_{\\textrm{eff}}${\\textless}/span{\\textgreater}] between locations that follow the emission sources and stabilizing compound concentrations for that region. Overall, the NPM allows researchers to easily model nucleation across diverse environments and estimate the concentration of non-sulfuric acid precursors using a condensation particle counter.{\\textless}/p{\\textgreater}},\n\tlanguage = {English},\n\tnumber = {12},\n\turldate = {2022-06-27},\n\tjournal = {Atmospheric Chemistry and Physics},\n\tauthor = {Johnson, Jack S. and Jen, Coty N.},\n\tmonth = jun,\n\tyear = {2022},\n\tnote = {Publisher: Copernicus GmbH},\n\tpages = {8287--8297},\n}\n\n
\n
\n\n\n
\n \\textlessp\\textgreater\\textlessstrong class=\"journal-contentHeaderColor\"\\textgreaterAbstract.\\textless/strong\\textgreater Observations over the last decade have demonstrated that the atmosphere contains potentially hundreds of compounds that can react with sulfuric acid to nucleate stable aerosol particles. Consequently, modeling atmospheric nucleation requires detailed knowledge of nucleation reaction kinetics and spatially and temporally resolved measurements of numerous precursor compounds. This study introduces the Nucleation Potential Model (NPM), a novel nucleation model that dramatically simplifies the diverse reactions between sulfuric acid and any combination of precursor gases. The NPM predicts 1 nm nucleation rates from only two measurable gas concentrations, regardless of whether all precursor gases are known. The NPM describes sulfuric acid nucleating with a parameterized base compound at an effective base concentration, [\\textlessspan class=\"inline-formula\"\\textgreaterB$_{\\textrm{eff}}$\\textless/span\\textgreater]. [\\textlessspan class=\"inline-formula\"\\textgreaterB$_{\\textrm{eff}}$\\textless/span\\textgreater] captures the ability of a compound or mixture to form stable clusters with sulfuric acid and is estimated from measured 1 nm particle concentrations. The NPM is applied to experimental and field observations of sulfuric acid nucleation to demonstrate how [\\textlessspan class=\"inline-formula\"\\textgreaterB$_{\\textrm{eff}}$\\textless/span\\textgreater] varies for different stabilizing compounds, mixtures, and sampling locations. Analysis of previous field observations shows distinct differences in [\\textlessspan class=\"inline-formula\"\\textgreaterB$_{\\textrm{eff}}$\\textless/span\\textgreater] between locations that follow the emission sources and stabilizing compound concentrations for that region. Overall, the NPM allows researchers to easily model nucleation across diverse environments and estimate the concentration of non-sulfuric acid precursors using a condensation particle counter.\\textless/p\\textgreater\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Aging of Volatile Organic Compounds in October 2017 Northern California Wildfire Plumes.\n \n \n \n \n\n\n \n Liang, Y.; Weber, R. J.; Misztal, P. K.; Jen, C. N.; and Goldstein, A. H.\n\n\n \n\n\n\n Environmental Science & Technology, 56(3): 1557–1567. February 2022.\n Publisher: American Chemical Society\n\n\n\n
\n\n\n\n \n \n \"AgingPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n  \n \n 1 download\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{liang_aging_2022,\n\ttitle = {Aging of {Volatile} {Organic} {Compounds} in {October} 2017 {Northern} {California} {Wildfire} {Plumes}},\n\tvolume = {56},\n\tissn = {0013-936X},\n\turl = {https://doi.org/10.1021/acs.est.1c05684},\n\tdoi = {10.1021/acs.est.1c05684},\n\tabstract = {In the western United States, the number and severity of large wildfires have been growing for decades. Biomass burning (BB) is a major source of volatile organic compounds (VOCs) to the atmosphere both globally and regionally. Following emission, BB VOCs are oxidized while being transported downwind, producing ozone, secondary organic aerosols, and secondary hazardous VOCs. In this research, we measured VOCs using proton transfer reaction time-of-flight mass spectrometry (PTR-ToF-MS) in an urban area 55–65 km downwind of the October 2017 Northern California wildfires. Nonaromatic oxygenated compounds were the dominant component of BB VOCs measured. In the smoke plumes, the VOCs account for 70–75\\% of the total observed organic carbon, with the remainder being particulate matter (with a diameter of {\\textless}2.5 μm, PM2.5). We show that the correlation of VOCs with furan (primary BB VOC) and maleic anhydride (secondary BB VOC) can indicate the origin of the VOCs. This was further confirmed by the diurnal variations of the VOCs and their concentration-weighted trajectories. Oxidation during transport consumed highly reactive compounds including benzenoids, furanoids, and terpenoids and produced more oxygenated VOCs. Furthermore, wildfire VOCs altered the ozone formation regime and raised the O3 levels in the San Francisco Bay Area.},\n\tnumber = {3},\n\turldate = {2022-03-18},\n\tjournal = {Environmental Science \\& Technology},\n\tauthor = {Liang, Yutong and Weber, Robert J. and Misztal, Pawel K. and Jen, Coty N. and Goldstein, Allen H.},\n\tmonth = feb,\n\tyear = {2022},\n\tnote = {Publisher: American Chemical Society},\n\tpages = {1557--1567},\n}\n\n
\n
\n\n\n
\n In the western United States, the number and severity of large wildfires have been growing for decades. Biomass burning (BB) is a major source of volatile organic compounds (VOCs) to the atmosphere both globally and regionally. Following emission, BB VOCs are oxidized while being transported downwind, producing ozone, secondary organic aerosols, and secondary hazardous VOCs. In this research, we measured VOCs using proton transfer reaction time-of-flight mass spectrometry (PTR-ToF-MS) in an urban area 55–65 km downwind of the October 2017 Northern California wildfires. Nonaromatic oxygenated compounds were the dominant component of BB VOCs measured. In the smoke plumes, the VOCs account for 70–75% of the total observed organic carbon, with the remainder being particulate matter (with a diameter of \\textless2.5 μm, PM2.5). We show that the correlation of VOCs with furan (primary BB VOC) and maleic anhydride (secondary BB VOC) can indicate the origin of the VOCs. This was further confirmed by the diurnal variations of the VOCs and their concentration-weighted trajectories. Oxidation during transport consumed highly reactive compounds including benzenoids, furanoids, and terpenoids and produced more oxygenated VOCs. Furthermore, wildfire VOCs altered the ozone formation regime and raised the O3 levels in the San Francisco Bay Area.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2021\n \n \n (2)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Chemical composition of PM$_{\\textrm{2.5}}$ in October 2017 Northern California wildfire plumes.\n \n \n \n \n\n\n \n Liang, Y.; Jen, C. N.; Weber, R. J.; Misztal, P. K.; and Goldstein, A. H.\n\n\n \n\n\n\n Atmospheric Chemistry and Physics, 21(7): 5719–5737. April 2021.\n Publisher: Copernicus GmbH\n\n\n\n
\n\n\n\n \n \n \"ChemicalPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n  \n \n 1 download\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{liang_chemical_2021,\n\ttitle = {Chemical composition of {PM}$_{\\textrm{2.5}}$ in {October} 2017 {Northern} {California} wildfire plumes},\n\tvolume = {21},\n\tcopyright = {All rights reserved},\n\tissn = {1680-7316},\n\turl = {https://acp.copernicus.org/articles/21/5719/2021/},\n\tdoi = {10.5194/acp-21-5719-2021},\n\tabstract = {{\\textless}p{\\textgreater}{\\textless}strong class="journal-contentHeaderColor"{\\textgreater}Abstract.{\\textless}/strong{\\textgreater} Wildfires have become more common and intense in the western US over recent decades due to a combination of historical land management practices and warming climate. Emissions from large-scale fires now frequently affect populated regions such as the San Francisco Bay Area during the fall wildfire season, with documented impacts of the resulting particulate matter on human health. Health impacts of exposure to wildfire emissions depend on the chemical composition of particulate matter, but the molecular composition of the real biomass burning organic aerosol (BBOA) that reaches large population centers remains insufficiently characterized. We took PM{\\textless}span class="inline-formula"{\\textgreater}$_{\\textrm{2.5}}${\\textless}/span{\\textgreater} (particles having aerodynamic diameters less than or equal to 2.5 {\\textless}span class="inline-formula"{\\textgreater}µm{\\textless}/span{\\textgreater}) samples at the University of California, Berkeley campus ({\\textless}span class="inline-formula"{\\textgreater}∼{\\textless}/span{\\textgreater} 60 km downwind of the fires) during the October 2017 Northern California wildfires period and analyzed molecular composition of OA using a two-dimensional gas chromatography coupled with high-resolution time-of-flight mass spectrometry (GC{\\textless}span class="inline-formula"{\\textgreater}×{\\textless}/span{\\textgreater}GC HR-ToF-MS). Sugar-like compounds were the most abundant component of BBOA, followed by mono-carboxylic acids, aromatic compounds, other oxygenated compounds, and terpenoids. The vast majority of compounds detected in smoke have unknown health impacts.{\\textless}/p{\\textgreater} {\\textless}p{\\textgreater}Regression models were trained to predict the saturation vapor pressure and averaged carbon oxidation state ({\\textless}span class="inline-formula"{\\textgreater}{\\textless}math xmlns="http://www.w3.org/1998/Math/MathML" id="M7" display="inline" overflow="scroll" dspmath="mathml"{\\textgreater}{\\textless}mover accent="true"{\\textgreater}{\\textless}mrow{\\textgreater}{\\textless}msub{\\textgreater}{\\textless}mi mathvariant="normal"{\\textgreater}OS{\\textless}/mi{\\textgreater}{\\textless}mi mathvariant="normal"{\\textgreater}c{\\textless}/mi{\\textgreater}{\\textless}/msub{\\textgreater}{\\textless}/mrow{\\textgreater}{\\textless}mo mathvariant="normal"{\\textgreater}‾{\\textless}/mo{\\textgreater}{\\textless}/mover{\\textgreater}{\\textless}/math{\\textgreater}{\\textless}span{\\textgreater}{\\textless}svg:svg xmlns:svg="http://www.w3.org/2000/svg" width="23pt" height="16pt" class="svg-formula" dspmath="mathimg" md5hash="2ed822bb6f358924dcfc775b7e6ab894"{\\textgreater}{\\textless}svg:image xmlns:xlink="http://www.w3.org/1999/xlink" xlink:href="acp-21-5719-2021-ie00001.svg" width="23pt" height="16pt" src="acp-21-5719-2021-ie00001.png"/{\\textgreater}{\\textless}/svg:svg{\\textgreater}{\\textless}/span{\\textgreater}{\\textless}/span{\\textgreater}) of detected compounds. The compounds speciated have a wide volatility distribution and most of them are highly oxygenated. In addition, time series of primary BBOA tracers observed in Berkeley were found to be indicative of the types of plants in the ecosystems burned in Napa and Sonoma, and could be used to differentiate the regions from which the smoke must have originated. Commonly used secondary BBOA markers like 4-nitrocatechol were enhanced when plumes aged, but their very fast formation caused them to have similar temporal variation as primary BBOA tracers. Using hierarchical clustering analysis, we classified compounds into seven factors indicative of their sources and transformation processes, identifying a unique daytime secondary BBOA factor. Chemicals associated with this factor include multifunctional acids and oxygenated aromatic compounds. These compounds have high {\\textless}span class="inline-formula"{\\textgreater}{\\textless}math xmlns="http://www.w3.org/1998/Math/MathML" id="M8" display="inline" overflow="scroll" dspmath="mathml"{\\textgreater}{\\textless}mover accent="true"{\\textgreater}{\\textless}mrow{\\textgreater}{\\textless}msub{\\textgreater}{\\textless}mi mathvariant="normal"{\\textgreater}OS{\\textless}/mi{\\textgreater}{\\textless}mi mathvariant="normal"{\\textgreater}c{\\textless}/mi{\\textgreater}{\\textless}/msub{\\textgreater}{\\textless}/mrow{\\textgreater}{\\textless}mo mathvariant="normal"{\\textgreater}‾{\\textless}/mo{\\textgreater}{\\textless}/mover{\\textgreater}{\\textless}/math{\\textgreater}{\\textless}span{\\textgreater}{\\textless}svg:svg xmlns:svg="http://www.w3.org/2000/svg" width="23pt" height="16pt" class="svg-formula" dspmath="mathimg" md5hash="4dc6696907890d1e0d046bd6b7b98d55"{\\textgreater}{\\textless}svg:image xmlns:xlink="http://www.w3.org/1999/xlink" xlink:href="acp-21-5719-2021-ie00002.svg" width="23pt" height="16pt" src="acp-21-5719-2021-ie00002.png"/{\\textgreater}{\\textless}/svg:svg{\\textgreater}{\\textless}/span{\\textgreater}{\\textless}/span{\\textgreater}, and they are also semi-volatile. We observed no net particle-phase organic carbon formation, which indicates an approximate balance between the mass of evaporated organic carbonaceous compounds and the addition of secondary organic carbonaceous compounds.{\\textless}/p{\\textgreater}},\n\tlanguage = {English},\n\tnumber = {7},\n\turldate = {2021-05-07},\n\tjournal = {Atmospheric Chemistry and Physics},\n\tauthor = {Liang, Yutong and Jen, Coty N. and Weber, Robert J. and Misztal, Pawel K. and Goldstein, Allen H.},\n\tmonth = apr,\n\tyear = {2021},\n\tnote = {Publisher: Copernicus GmbH},\n\tpages = {5719--5737},\n}\n\n
\n
\n\n\n
\n \\textlessp\\textgreater\\textlessstrong class=\"journal-contentHeaderColor\"\\textgreaterAbstract.\\textless/strong\\textgreater Wildfires have become more common and intense in the western US over recent decades due to a combination of historical land management practices and warming climate. Emissions from large-scale fires now frequently affect populated regions such as the San Francisco Bay Area during the fall wildfire season, with documented impacts of the resulting particulate matter on human health. Health impacts of exposure to wildfire emissions depend on the chemical composition of particulate matter, but the molecular composition of the real biomass burning organic aerosol (BBOA) that reaches large population centers remains insufficiently characterized. We took PM\\textlessspan class=\"inline-formula\"\\textgreater$_{\\textrm{2.5}}$\\textless/span\\textgreater (particles having aerodynamic diameters less than or equal to 2.5 \\textlessspan class=\"inline-formula\"\\textgreaterµm\\textless/span\\textgreater) samples at the University of California, Berkeley campus (\\textlessspan class=\"inline-formula\"\\textgreater∼\\textless/span\\textgreater 60 km downwind of the fires) during the October 2017 Northern California wildfires period and analyzed molecular composition of OA using a two-dimensional gas chromatography coupled with high-resolution time-of-flight mass spectrometry (GC\\textlessspan class=\"inline-formula\"\\textgreater×\\textless/span\\textgreaterGC HR-ToF-MS). Sugar-like compounds were the most abundant component of BBOA, followed by mono-carboxylic acids, aromatic compounds, other oxygenated compounds, and terpenoids. The vast majority of compounds detected in smoke have unknown health impacts.\\textless/p\\textgreater \\textlessp\\textgreaterRegression models were trained to predict the saturation vapor pressure and averaged carbon oxidation state (\\textlessspan class=\"inline-formula\"\\textgreater\\textlessmath xmlns=\"http://www.w3.org/1998/Math/MathML\" id=\"M7\" display=\"inline\" overflow=\"scroll\" dspmath=\"mathml\"\\textgreater\\textlessmover accent=\"true\"\\textgreater\\textlessmrow\\textgreater\\textlessmsub\\textgreater\\textlessmi mathvariant=\"normal\"\\textgreaterOS\\textless/mi\\textgreater\\textlessmi mathvariant=\"normal\"\\textgreaterc\\textless/mi\\textgreater\\textless/msub\\textgreater\\textless/mrow\\textgreater\\textlessmo mathvariant=\"normal\"\\textgreater‾\\textless/mo\\textgreater\\textless/mover\\textgreater\\textless/math\\textgreater\\textlessspan\\textgreater\\textlesssvg:svg xmlns:svg=\"http://www.w3.org/2000/svg\" width=\"23pt\" height=\"16pt\" class=\"svg-formula\" dspmath=\"mathimg\" md5hash=\"2ed822bb6f358924dcfc775b7e6ab894\"\\textgreater\\textlesssvg:image xmlns:xlink=\"http://www.w3.org/1999/xlink\" xlink:href=\"acp-21-5719-2021-ie00001.svg\" width=\"23pt\" height=\"16pt\" src=\"acp-21-5719-2021-ie00001.png\"/\\textgreater\\textless/svg:svg\\textgreater\\textless/span\\textgreater\\textless/span\\textgreater) of detected compounds. The compounds speciated have a wide volatility distribution and most of them are highly oxygenated. In addition, time series of primary BBOA tracers observed in Berkeley were found to be indicative of the types of plants in the ecosystems burned in Napa and Sonoma, and could be used to differentiate the regions from which the smoke must have originated. Commonly used secondary BBOA markers like 4-nitrocatechol were enhanced when plumes aged, but their very fast formation caused them to have similar temporal variation as primary BBOA tracers. Using hierarchical clustering analysis, we classified compounds into seven factors indicative of their sources and transformation processes, identifying a unique daytime secondary BBOA factor. Chemicals associated with this factor include multifunctional acids and oxygenated aromatic compounds. These compounds have high \\textlessspan class=\"inline-formula\"\\textgreater\\textlessmath xmlns=\"http://www.w3.org/1998/Math/MathML\" id=\"M8\" display=\"inline\" overflow=\"scroll\" dspmath=\"mathml\"\\textgreater\\textlessmover accent=\"true\"\\textgreater\\textlessmrow\\textgreater\\textlessmsub\\textgreater\\textlessmi mathvariant=\"normal\"\\textgreaterOS\\textless/mi\\textgreater\\textlessmi mathvariant=\"normal\"\\textgreaterc\\textless/mi\\textgreater\\textless/msub\\textgreater\\textless/mrow\\textgreater\\textlessmo mathvariant=\"normal\"\\textgreater‾\\textless/mo\\textgreater\\textless/mover\\textgreater\\textless/math\\textgreater\\textlessspan\\textgreater\\textlesssvg:svg xmlns:svg=\"http://www.w3.org/2000/svg\" width=\"23pt\" height=\"16pt\" class=\"svg-formula\" dspmath=\"mathimg\" md5hash=\"4dc6696907890d1e0d046bd6b7b98d55\"\\textgreater\\textlesssvg:image xmlns:xlink=\"http://www.w3.org/1999/xlink\" xlink:href=\"acp-21-5719-2021-ie00002.svg\" width=\"23pt\" height=\"16pt\" src=\"acp-21-5719-2021-ie00002.png\"/\\textgreater\\textless/svg:svg\\textgreater\\textless/span\\textgreater\\textless/span\\textgreater, and they are also semi-volatile. We observed no net particle-phase organic carbon formation, which indicates an approximate balance between the mass of evaporated organic carbonaceous compounds and the addition of secondary organic carbonaceous compounds.\\textless/p\\textgreater\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n A tutorial guide on new particle formation experiments using a laminar flow reactor.\n \n \n \n \n\n\n \n Fomete, S. K. W.; Johnson, J. S.; Casalnuovo, D.; and Jen, C. N.\n\n\n \n\n\n\n Journal of Aerosol Science, 157: 105808. September 2021.\n \n\n\n\n
\n\n\n\n \n \n \"APaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n  \n \n 1 download\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{fomete_tutorial_2021,\n\ttitle = {A tutorial guide on new particle formation experiments using a laminar flow reactor},\n\tvolume = {157},\n\tcopyright = {All rights reserved},\n\tissn = {0021-8502},\n\turl = {https://www.sciencedirect.com/science/article/pii/S0021850221005395},\n\tdoi = {10.1016/j.jaerosci.2021.105808},\n\tabstract = {New particle formation (NPF) produces about 50\\% of the global cloud condensation nuclei in the troposphere. As such, NPF plays a crucial role in climate. Despite advancements in instrumentation capable of measuring freshly formed aerosol particles down to {\\textasciitilde} 1 nm in diameter, the mechanisms behind NPF remain understudied due to the complex composition and chemistry of the atmosphere. Nucleation is the first step of NPF and involves gaseous precursors reacting to form stable clusters; consequently, it is essential to understand the reaction kinetics behind nucleation reactions. Controlled laboratory experiments have previously been used to examine these reactions, which can occur at extremely low reactant concentrations (i.e., parts per quadrillion level, 105 cm−3) or lower. Nucleation experiments require pristine conditions for the reactions to proceed without interference from unpredictable contaminants. Here, a low-cost flow reactor is presented that minimizes contamination and allows for nucleation kinetics to be observed. The layout and setup of an example reactor are presented with a brief discussion on how to operate the reactor to ensure cleanliness and repeatability. In addition, methods for quantifying nucleation reactants as well as analytical measurement techniques to adequately measure nucleation kinetics in this flow reactor system are described. This experimental protocol can be employed to characterize nucleation reactions that can ultimately be used to develop nucleation models important for predicting how aerosol particles influence climate.},\n\tlanguage = {en},\n\turldate = {2021-05-16},\n\tjournal = {Journal of Aerosol Science},\n\tauthor = {Fomete, Sandra K. W. and Johnson, Jack S. and Casalnuovo, Dominic and Jen, Coty N.},\n\tmonth = sep,\n\tyear = {2021},\n\tkeywords = {New particle formation (NPF), Nucleation, Nucleation flow reactor},\n\tpages = {105808},\n}\n\n
\n
\n\n\n
\n New particle formation (NPF) produces about 50% of the global cloud condensation nuclei in the troposphere. As such, NPF plays a crucial role in climate. Despite advancements in instrumentation capable of measuring freshly formed aerosol particles down to ~ 1 nm in diameter, the mechanisms behind NPF remain understudied due to the complex composition and chemistry of the atmosphere. Nucleation is the first step of NPF and involves gaseous precursors reacting to form stable clusters; consequently, it is essential to understand the reaction kinetics behind nucleation reactions. Controlled laboratory experiments have previously been used to examine these reactions, which can occur at extremely low reactant concentrations (i.e., parts per quadrillion level, 105 cm−3) or lower. Nucleation experiments require pristine conditions for the reactions to proceed without interference from unpredictable contaminants. Here, a low-cost flow reactor is presented that minimizes contamination and allows for nucleation kinetics to be observed. The layout and setup of an example reactor are presented with a brief discussion on how to operate the reactor to ensure cleanliness and repeatability. In addition, methods for quantifying nucleation reactants as well as analytical measurement techniques to adequately measure nucleation kinetics in this flow reactor system are described. This experimental protocol can be employed to characterize nucleation reactions that can ultimately be used to develop nucleation models important for predicting how aerosol particles influence climate.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2019\n \n \n (1)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Speciated and total emission factors of particulate organics from burning western US wildland fuels and their dependence on combustion efficiency.\n \n \n \n \n\n\n \n Jen, C. N.; Hatch, L. E.; Selimovic, V.; Yokelson, R. J.; Weber, R.; Fernandez, A. E.; Kreisberg, N. M.; Barsanti, K. C.; and Goldstein, A. H.\n\n\n \n\n\n\n Atmospheric Chemistry and Physics, 19(2): 1013–1026. January 2019.\n \n\n\n\n
\n\n\n\n \n \n \"SpeciatedPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{jen_speciated_2019,\n\ttitle = {Speciated and total emission factors of particulate organics from burning western {US} wildland fuels and their dependence on combustion efficiency},\n\tvolume = {19},\n\tcopyright = {All rights reserved},\n\tissn = {1680-7316},\n\turl = {https://www.atmos-chem-phys.net/19/1013/2019/},\n\tdoi = {https://doi.org/10.5194/acp-19-1013-2019},\n\tabstract = {{\\textless}p{\\textgreater}{\\textless}strong{\\textgreater}Abstract.{\\textless}/strong{\\textgreater} Western US wildlands experience frequent and large-scale wildfires which are predicted to increase in the future. As a result, wildfire smoke emissions are expected to play an increasing role in atmospheric chemistry while negatively impacting regional air quality and human health. Understanding the impacts of smoke on the environment is informed by identifying and quantifying the chemical compounds that are emitted during wildfires and by providing empirical relationships that describe how the amount and composition of the emissions change based upon different fire conditions and fuels. This study examined particulate organic compounds emitted from burning common western US wildland fuels at the US Forest Service Fire Science Laboratory. Thousands of intermediate and semi-volatile organic compounds (I/SVOCs) were separated and quantified into fire-integrated emission factors (EFs) using a thermal desorption, two-dimensional gas chromatograph with online derivatization coupled to an electron ionization/vacuum ultraviolet high-resolution time-of-flight mass spectrometer (TD-GC{\\textless}span class="thinspace"{\\textgreater}{\\textless}/span{\\textgreater}{\\textless}span class="inline-formula"{\\textgreater}×{\\textless}/span{\\textgreater}{\\textless}span class="thinspace"{\\textgreater}{\\textless}/span{\\textgreater}GC-EI/VUV-HRToFMS). Mass spectra, EFs as a function of modified combustion efficiency (MCE), fuel source, and other defining characteristics for the separated compounds are provided in the accompanying mass spectral library. Results show that EFs for total organic carbon (OC), chemical families of I/SVOCs, and most individual I/SVOCs span 2–5 orders of magnitude, with higher EFs at smoldering conditions (low MCE) than flaming. Logarithmic fits applied to the observations showed that log (EFs) for particulate organic compounds were inversely proportional to MCE. These measurements and relationships provide useful estimates of EFs for OC, elemental carbon (EC), organic chemical families, and individual I/SVOCs as a function of fire conditions.{\\textless}/p{\\textgreater}},\n\tlanguage = {English},\n\tnumber = {2},\n\turldate = {2019-01-25},\n\tjournal = {Atmospheric Chemistry and Physics},\n\tauthor = {Jen, Coty N. and Hatch, Lindsay E. and Selimovic, Vanessa and Yokelson, Robert J. and Weber, Robert and Fernandez, Arantza E. and Kreisberg, Nathan M. and Barsanti, Kelley C. and Goldstein, Allen H.},\n\tmonth = jan,\n\tyear = {2019},\n\tpages = {1013--1026},\n}\n\n
\n
\n\n\n
\n \\textlessp\\textgreater\\textlessstrong\\textgreaterAbstract.\\textless/strong\\textgreater Western US wildlands experience frequent and large-scale wildfires which are predicted to increase in the future. As a result, wildfire smoke emissions are expected to play an increasing role in atmospheric chemistry while negatively impacting regional air quality and human health. Understanding the impacts of smoke on the environment is informed by identifying and quantifying the chemical compounds that are emitted during wildfires and by providing empirical relationships that describe how the amount and composition of the emissions change based upon different fire conditions and fuels. This study examined particulate organic compounds emitted from burning common western US wildland fuels at the US Forest Service Fire Science Laboratory. Thousands of intermediate and semi-volatile organic compounds (I/SVOCs) were separated and quantified into fire-integrated emission factors (EFs) using a thermal desorption, two-dimensional gas chromatograph with online derivatization coupled to an electron ionization/vacuum ultraviolet high-resolution time-of-flight mass spectrometer (TD-GC\\textlessspan class=\"thinspace\"\\textgreater\\textless/span\\textgreater\\textlessspan class=\"inline-formula\"\\textgreater×\\textless/span\\textgreater\\textlessspan class=\"thinspace\"\\textgreater\\textless/span\\textgreaterGC-EI/VUV-HRToFMS). Mass spectra, EFs as a function of modified combustion efficiency (MCE), fuel source, and other defining characteristics for the separated compounds are provided in the accompanying mass spectral library. Results show that EFs for total organic carbon (OC), chemical families of I/SVOCs, and most individual I/SVOCs span 2–5 orders of magnitude, with higher EFs at smoldering conditions (low MCE) than flaming. Logarithmic fits applied to the observations showed that log (EFs) for particulate organic compounds were inversely proportional to MCE. These measurements and relationships provide useful estimates of EFs for OC, elemental carbon (EC), organic chemical families, and individual I/SVOCs as a function of fire conditions.\\textless/p\\textgreater\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2018\n \n \n (3)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n High Hydroquinone Emissions from Burning Manzanita.\n \n \n \n \n\n\n \n Jen, C. N.; Liang, Y.; Hatch, L. E.; Kreisberg, N. M.; Stamatis, C.; Kristensen, K.; Battles, J. J.; Stephens, S. L.; York, R. A.; Barsanti, K. C.; and Goldstein, A. H.\n\n\n \n\n\n\n Environmental Science & Technology Letters, 5(6): 309–314. May 2018.\n \n\n\n\n
\n\n\n\n \n \n \"HighPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n  \n \n 1 download\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{jen_high_2018,\n\ttitle = {High {Hydroquinone} {Emissions} from {Burning} {Manzanita}},\n\tvolume = {5},\n\tcopyright = {All rights reserved},\n\turl = {https://doi.org/10.1021/acs.estlett.8b00222},\n\tdoi = {10.1021/acs.estlett.8b00222},\n\tabstract = {California wildfires are becoming larger and more frequent because of climate change and historical fire suppression. The 2017 fire season was record-breaking in terms of monetary damage, area burned, and human casualties. In addition, roughly 20 million people were exposed to dense wildfire smoke for days. Understanding the health impacts of wildfire smoke requires detailed chemical speciation of smoke produced from different fuels. This study demonstrates the unique chemical fingerprint observed in smoke from burning manzanita, a common chaparral and forest understory shrub found in several ecosystems of California. Burning manzanita during the FIREX Fire Laboratory experiments emitted hydroquinone (1,4-dihydroxybenzene with an emission factor of 0.4 g/kg) and two sterol/triterpenoid tracer compounds at levels up to 100 times higher than those of the other common wildland fuels in California such as pine trees, other shrubs, grasses, and duff. Additionally, these compounds were detected in Berkeley, CA, from smoke produced during the October 2017 wildfires in northern California, a region where manzanita grows. In contrast, the identified fingerprint for manzanita burning emissions was not observed during prescribed fires of a mixed conifer forest in California’s Sierra Nevada, indicating negligible amounts of manzanita were burned. As confirmed by shrub inventory data collected prior to the burns, small amounts of manzanita remain after prescribed burning, a low-severity forest management technique, but larger amounts can occur after recovery from high-severity events like wildfires. Results from this study show that chemical signatures in smoke can be traced back to specific fuels like manzanita and that forest management techniques can be used to limit certain types of wildfire emissions.},\n\tnumber = {6},\n\turldate = {2018-05-16},\n\tjournal = {Environmental Science \\& Technology Letters},\n\tauthor = {Jen, Coty N. and Liang, Yutong and Hatch, Lindsay E. and Kreisberg, Nathan M. and Stamatis, Christos and Kristensen, Kasper and Battles, John J. and Stephens, Scott L. and York, Robert A. and Barsanti, Kelley C. and Goldstein, Allen H.},\n\tmonth = may,\n\tyear = {2018},\n\tpages = {309--314},\n}\n\n
\n
\n\n\n
\n California wildfires are becoming larger and more frequent because of climate change and historical fire suppression. The 2017 fire season was record-breaking in terms of monetary damage, area burned, and human casualties. In addition, roughly 20 million people were exposed to dense wildfire smoke for days. Understanding the health impacts of wildfire smoke requires detailed chemical speciation of smoke produced from different fuels. This study demonstrates the unique chemical fingerprint observed in smoke from burning manzanita, a common chaparral and forest understory shrub found in several ecosystems of California. Burning manzanita during the FIREX Fire Laboratory experiments emitted hydroquinone (1,4-dihydroxybenzene with an emission factor of 0.4 g/kg) and two sterol/triterpenoid tracer compounds at levels up to 100 times higher than those of the other common wildland fuels in California such as pine trees, other shrubs, grasses, and duff. Additionally, these compounds were detected in Berkeley, CA, from smoke produced during the October 2017 wildfires in northern California, a region where manzanita grows. In contrast, the identified fingerprint for manzanita burning emissions was not observed during prescribed fires of a mixed conifer forest in California’s Sierra Nevada, indicating negligible amounts of manzanita were burned. As confirmed by shrub inventory data collected prior to the burns, small amounts of manzanita remain after prescribed burning, a low-severity forest management technique, but larger amounts can occur after recovery from high-severity events like wildfires. Results from this study show that chemical signatures in smoke can be traced back to specific fuels like manzanita and that forest management techniques can be used to limit certain types of wildfire emissions.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Evolution of the chemical fingerprint of biomass burning organic aerosol during aging.\n \n \n \n \n\n\n \n Bertrand, A.; Stefenelli, G.; Jen, C. N.; Pieber, S. M.; Bruns, E. A.; Ni, H.; Temime-Roussel, B.; Slowik, J. G.; Goldstein, A. H.; Haddad, I. E.; Baltensperger, U.; Prévôt, A. S. H.; Wortham, H.; and Marchand, N.\n\n\n \n\n\n\n Atmospheric Chemistry and Physics, 18(10): 7607–7624. June 2018.\n \n\n\n\n
\n\n\n\n \n \n \"EvolutionPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{bertrand_evolution_2018,\n\ttitle = {Evolution of the chemical fingerprint of biomass burning organic aerosol during aging},\n\tvolume = {18},\n\tcopyright = {All rights reserved},\n\tissn = {1680-7316},\n\turl = {https://www.atmos-chem-phys.net/18/7607/2018/acp-18-7607-2018.html},\n\tdoi = {https://doi.org/10.5194/acp-18-7607-2018},\n\tabstract = {{\\textless}p{\\textgreater}{\\textless}strong{\\textgreater}Abstract.{\\textless}/strong{\\textgreater} A thermal desorption aerosol gas chromatograph coupled to a high resolution \\&ndash; time of flight \\&ndash; aerosol mass spectrometer (TAG-AMS) was connected to an atmospheric chamber for the molecular characterization of the evolution of organic aerosol (OA) emitted by woodstove appliances for residential heating. Two log woodstoves (old and modern) and one pellet stove were operated under typical conditions. Emissions were aged during a time equivalent to 5{\\textless}span class="thinspace"{\\textgreater}{\\textless}/span{\\textgreater}h of atmospheric aging. The five to seven samples were collected and analyzed with the TAG-AMS during each experiment. We detected and quantified over 70 compounds, including levoglucosan and nitrocatechols. We calculate the emission factor (EF) of these tracers in the primary emissions and highlight the influence of the combustion efficiency on these emissions. Smoldering combustion contributes to a higher EF and a more complex composition. We also demonstrate the effect of atmospheric aging on the chemical fingerprint. The tracers are sorted into three categories according to the evolution of their concentration: primary compounds, non-conventional primary compounds, and secondary compounds. For each, we provide a quantitative overview of their contribution to the OA mass at different times of the photo-oxidative process.{\\textless}/p{\\textgreater}},\n\tlanguage = {English},\n\tnumber = {10},\n\turldate = {2019-04-30},\n\tjournal = {Atmospheric Chemistry and Physics},\n\tauthor = {Bertrand, Amelie and Stefenelli, Giulia and Jen, Coty N. and Pieber, Simone M. and Bruns, Emily A. and Ni, Haiyan and Temime-Roussel, Brice and Slowik, Jay G. and Goldstein, Allen H. and Haddad, Imad El and Baltensperger, Urs and Prévôt, André S. H. and Wortham, Henri and Marchand, Nicolas},\n\tmonth = jun,\n\tyear = {2018},\n\tpages = {7607--7624},\n}\n\n
\n
\n\n\n
\n \\textlessp\\textgreater\\textlessstrong\\textgreaterAbstract.\\textless/strong\\textgreater A thermal desorption aerosol gas chromatograph coupled to a high resolution – time of flight – aerosol mass spectrometer (TAG-AMS) was connected to an atmospheric chamber for the molecular characterization of the evolution of organic aerosol (OA) emitted by woodstove appliances for residential heating. Two log woodstoves (old and modern) and one pellet stove were operated under typical conditions. Emissions were aged during a time equivalent to 5\\textlessspan class=\"thinspace\"\\textgreater\\textless/span\\textgreaterh of atmospheric aging. The five to seven samples were collected and analyzed with the TAG-AMS during each experiment. We detected and quantified over 70 compounds, including levoglucosan and nitrocatechols. We calculate the emission factor (EF) of these tracers in the primary emissions and highlight the influence of the combustion efficiency on these emissions. Smoldering combustion contributes to a higher EF and a more complex composition. We also demonstrate the effect of atmospheric aging on the chemical fingerprint. The tracers are sorted into three categories according to the evolution of their concentration: primary compounds, non-conventional primary compounds, and secondary compounds. For each, we provide a quantitative overview of their contribution to the OA mass at different times of the photo-oxidative process.\\textless/p\\textgreater\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Measurements of I/SVOCs in biomass burning smoke using solid-phase extraction disks and two-dimensional gas chromatography.\n \n \n \n\n\n \n Hatch, L. E.; Rivas-Ubach, A.; Jen, C. N.; Lipton, M.; Goldstein, A. H.; and Barsanti, K. C.\n\n\n \n\n\n\n Atmos. Chem. Phys., (18): 17801–17817. 2018.\n \n\n\n\n
\n\n\n\n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{hatch_measurements_2018,\n\ttitle = {Measurements of {I}/{SVOCs} in biomass burning smoke using solid-phase extraction disks and two-dimensional gas chromatography},\n\tcopyright = {All rights reserved},\n\tdoi = {https://doi.org/10.5194/acp-18-17801-2018},\n\tnumber = {18},\n\tjournal = {Atmos. Chem. Phys.},\n\tauthor = {Hatch, Lindsay E. and Rivas-Ubach, Albert and Jen, Coty N. and Lipton, Mary and Goldstein, Allen H. and Barsanti, Kelley C.},\n\tyear = {2018},\n\tpages = {17801--17817},\n}\n\n
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2017\n \n \n (2)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n New particle formation from sulfuric acid and amines: Comparison of monomethylamine, dimethylamine, and trimethylamine.\n \n \n \n \n\n\n \n Olenius, T.; Halonen, R.; Kurtén, T.; Henschel, H.; Kupiainen-Määttä, O.; Ortega, I. K.; Jen, C. N.; Vehkamäki, H.; and Riipinen, I.\n\n\n \n\n\n\n Journal of Geophysical Research: Atmospheres, 122(13): 2017JD026501. July 2017.\n \n\n\n\n
\n\n\n\n \n \n \"NewPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{olenius_new_2017,\n\ttitle = {New particle formation from sulfuric acid and amines: {Comparison} of monomethylamine, dimethylamine, and trimethylamine},\n\tvolume = {122},\n\tcopyright = {All rights reserved},\n\tissn = {2169-8996},\n\tshorttitle = {New particle formation from sulfuric acid and amines},\n\turl = {http://onlinelibrary.wiley.com/doi/10.1002/2017JD026501/abstract},\n\tdoi = {10.1002/2017JD026501},\n\tabstract = {Amines are bases that originate from both anthropogenic and natural sources, and they are recognized as candidates to participate in atmospheric aerosol particle formation together with sulfuric acid. Monomethylamine, dimethylamine, and trimethylamine (MMA, DMA, and TMA, respectively) have been shown to enhance sulfuric acid-driven particle formation more efficiently than ammonia, but both theory and laboratory experiments suggest that there are differences in their enhancing potentials. However, as quantitative concentrations and thermochemical properties of different amines remain relatively uncertain, and also for computational reasons, the compounds have been treated as a single surrogate amine species in large-scale modeling studies. In this work, the differences and similarities of MMA, DMA, and TMA are studied by simulations of molecular cluster formation from sulfuric acid, water, and each of the three amines. Quantum chemistry-based cluster evaporation rate constants are applied in a cluster population dynamics model to yield cluster concentrations and formation rates at boundary layer conditions. While there are differences, for instance, in the clustering mechanisms and cluster hygroscopicity for the three amines, DMA and TMA can be approximated as a lumped species. Formation of nanometer-sized particles and its dependence on ambient conditions is roughly similar for these two: both efficiently form clusters with sulfuric acid, and cluster formation is rather insensitive to changes in temperature and relative humidity. Particle formation from sulfuric acid and MMA is weaker and significantly more sensitive to ambient conditions. Therefore, merging MMA together with DMA and TMA introduces inaccuracies in sulfuric acid-amine particle formation schemes.},\n\tlanguage = {en},\n\tnumber = {13},\n\turldate = {2018-01-23},\n\tjournal = {Journal of Geophysical Research: Atmospheres},\n\tauthor = {Olenius, Tinja and Halonen, Roope and Kurtén, Theo and Henschel, Henning and Kupiainen-Määttä, Oona and Ortega, Ismael K. and Jen, Coty N. and Vehkamäki, Hanna and Riipinen, Ilona},\n\tmonth = jul,\n\tyear = {2017},\n\tkeywords = {0305 Aerosols and particles, 0317 Chemical kinetic and photochemical properties, 3307 Boundary layer processes, 3367 Theoretical modeling, amines, atmospheric new particle formation, molecular cluster kinetics, particle formation rate, quantum chemistry, sulfuric acid},\n\tpages = {2017JD026501},\n}\n\n
\n
\n\n\n
\n Amines are bases that originate from both anthropogenic and natural sources, and they are recognized as candidates to participate in atmospheric aerosol particle formation together with sulfuric acid. Monomethylamine, dimethylamine, and trimethylamine (MMA, DMA, and TMA, respectively) have been shown to enhance sulfuric acid-driven particle formation more efficiently than ammonia, but both theory and laboratory experiments suggest that there are differences in their enhancing potentials. However, as quantitative concentrations and thermochemical properties of different amines remain relatively uncertain, and also for computational reasons, the compounds have been treated as a single surrogate amine species in large-scale modeling studies. In this work, the differences and similarities of MMA, DMA, and TMA are studied by simulations of molecular cluster formation from sulfuric acid, water, and each of the three amines. Quantum chemistry-based cluster evaporation rate constants are applied in a cluster population dynamics model to yield cluster concentrations and formation rates at boundary layer conditions. While there are differences, for instance, in the clustering mechanisms and cluster hygroscopicity for the three amines, DMA and TMA can be approximated as a lumped species. Formation of nanometer-sized particles and its dependence on ambient conditions is roughly similar for these two: both efficiently form clusters with sulfuric acid, and cluster formation is rather insensitive to changes in temperature and relative humidity. Particle formation from sulfuric acid and MMA is weaker and significantly more sensitive to ambient conditions. Therefore, merging MMA together with DMA and TMA introduces inaccuracies in sulfuric acid-amine particle formation schemes.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Computational Fluid Dynamics Studies of a Flow Reactor: Free Energies of Clusters of Sulfuric Acid with NH3 or Dimethyl Amine.\n \n \n \n \n\n\n \n Hanson, D. R.; Bier, I.; Panta, B.; Jen, C. N.; and McMurry, P. H.\n\n\n \n\n\n\n The Journal of Physical Chemistry A, 121(20): 3976–3990. May 2017.\n \n\n\n\n
\n\n\n\n \n \n \"ComputationalPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{hanson_computational_2017,\n\ttitle = {Computational {Fluid} {Dynamics} {Studies} of a {Flow} {Reactor}: {Free} {Energies} of {Clusters} of {Sulfuric} {Acid} with {NH3} or {Dimethyl} {Amine}},\n\tvolume = {121},\n\tcopyright = {All rights reserved},\n\tissn = {1089-5639},\n\tshorttitle = {Computational {Fluid} {Dynamics} {Studies} of a {Flow} {Reactor}},\n\turl = {http://dx.doi.org/10.1021/acs.jpca.7b00252},\n\tdoi = {10.1021/acs.jpca.7b00252},\n\tabstract = {Computational fluid dynamics simulations of a flow reactor provided 3D spatial distributions of its temperature and flow profiles and abundances of sulfuric acid, nitrogeneous base, and the acid–base clusters formed from them. Clusters were simulated via their kinetic formation and decomposition involving sulfuric acid and base molecules. Temperature and flow profiles and the base and sulfuric acid distributions are characterized and the latter is compared to mass spectrometer measurements. Concentrations of simulated clusters of sulfuric acid with either NH3 or dimethylamine were compared to experimentally measured particle concentrations. Cluster thermodynamics were adjusted to better the agreement between simulated and experimental results. Free energies of acid–base clusters derived here are also compared to recent quantum chemistry calculations. Sensitivities to the thermodynamics were explored with a 2D laminar flow simulation and the abundance of large clusters was most sensitive to the thermodynamics of the smallest cluster, consisting of 1 base and 1 acid. Comparisons of this model to the computational fluid dynamics models provide verification of the implemented cluster chemistry. A box model was used to calculate nucleation rates for the conditions of other experimental work, and to provide predictions of nucleation for typical atmospheric conditions.},\n\tnumber = {20},\n\turldate = {2018-01-23},\n\tjournal = {The Journal of Physical Chemistry A},\n\tauthor = {Hanson, D. R. and Bier, I. and Panta, B. and Jen, C. N. and McMurry, P. H.},\n\tmonth = may,\n\tyear = {2017},\n\tpages = {3976--3990},\n}\n\n
\n
\n\n\n
\n Computational fluid dynamics simulations of a flow reactor provided 3D spatial distributions of its temperature and flow profiles and abundances of sulfuric acid, nitrogeneous base, and the acid–base clusters formed from them. Clusters were simulated via their kinetic formation and decomposition involving sulfuric acid and base molecules. Temperature and flow profiles and the base and sulfuric acid distributions are characterized and the latter is compared to mass spectrometer measurements. Concentrations of simulated clusters of sulfuric acid with either NH3 or dimethylamine were compared to experimentally measured particle concentrations. Cluster thermodynamics were adjusted to better the agreement between simulated and experimental results. Free energies of acid–base clusters derived here are also compared to recent quantum chemistry calculations. Sensitivities to the thermodynamics were explored with a 2D laminar flow simulation and the abundance of large clusters was most sensitive to the thermodynamics of the smallest cluster, consisting of 1 base and 1 acid. Comparisons of this model to the computational fluid dynamics models provide verification of the implemented cluster chemistry. A box model was used to calculate nucleation rates for the conditions of other experimental work, and to provide predictions of nucleation for typical atmospheric conditions.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2016\n \n \n (4)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Multiple new-particle growth pathways observed at the US DOE Southern Great Plains field site.\n \n \n \n \n\n\n \n Hodshire, A. L.; Lawler, M. J.; Zhao, J.; Ortega, J.; Jen, C.; Yli-Juuti, T.; Brewer, J. F.; Kodros, J. K.; Barsanti, K. C.; Hanson, D. R.; McMurry, P. H.; Smith, J. N.; and Pierce, J. R.\n\n\n \n\n\n\n Atmospheric Chemistry and Physics, 16(14): 9321–9348. July 2016.\n \n\n\n\n
\n\n\n\n \n \n \"MultiplePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{hodshire_multiple_2016,\n\ttitle = {Multiple new-particle growth pathways observed at the {US} {DOE} {Southern} {Great} {Plains} field site},\n\tvolume = {16},\n\tcopyright = {All rights reserved},\n\tissn = {1680-7316},\n\turl = {https://www.atmos-chem-phys.net/16/9321/2016/acp-16-9321-2016.html},\n\tdoi = {https://doi.org/10.5194/acp-16-9321-2016},\n\tabstract = {{\\textless}p{\\textgreater}{\\textless}strong{\\textgreater}Abstract.{\\textless}/strong{\\textgreater} New-particle formation (NPF) is a significant source of aerosol particles into the atmosphere. However, these particles are initially too small to have climatic importance and must grow, primarily through net uptake of low-volatility species, from diameters ∼ {\\textless}span class="thinspace"{\\textgreater}{\\textless}/span{\\textgreater}1 to 30–100{\\textless}span class="thinspace"{\\textgreater}{\\textless}/span{\\textgreater}nm in order to potentially impact climate. There are currently uncertainties in the physical and chemical processes associated with the growth of these freshly formed particles that lead to uncertainties in aerosol-climate modeling. Four main pathways for new-particle growth have been identified: condensation of sulfuric-acid vapor (and associated bases when available), condensation of organic vapors, uptake of organic acids through acid–base chemistry in the particle phase, and accretion of organic molecules in the particle phase to create a lower-volatility compound that then contributes to the aerosol mass. The relative importance of each pathway is uncertain and is the focus of this work. {\\textless}br{\\textgreater}{\\textless}br{\\textgreater} The 2013 New Particle Formation Study (NPFS) measurement campaign took place at the DOE Southern Great Plains (SGP) facility in Lamont, Oklahoma, during spring 2013. Measured gas- and particle-phase compositions during these new-particle growth events suggest three distinct growth pathways: (1) growth by primarily organics, (2) growth by primarily sulfuric acid and ammonia, and (3) growth by primarily sulfuric acid and associated bases and organics. To supplement the measurements, we used the particle growth model MABNAG (Model for Acid–Base chemistry in NAnoparticle Growth) to gain further insight into the growth processes on these 3 days at SGP. MABNAG simulates growth from (1) sulfuric-acid condensation (and subsequent salt formation with ammonia or amines), (2) near-irreversible condensation from nonreactive extremely low-volatility organic compounds (ELVOCs), and (3) organic-acid condensation and subsequent salt formation with ammonia or amines. MABNAG is able to corroborate the observed differing growth pathways, while also predicting that ELVOCs contribute more to growth than organic salt formation. However, most MABNAG model simulations tend to underpredict the observed growth rates between 10 and 20{\\textless}span class="thinspace"{\\textgreater}{\\textless}/span{\\textgreater}nm in diameter; this underprediction may come from neglecting the contributions to growth from semi-to-low-volatility species or accretion reactions. Our results suggest that in addition to sulfuric acid, ELVOCs are also very important for growth in this rural setting. We discuss the limitations of our study that arise from not accounting for semi- and low-volatility organics, as well as nitrogen-containing species beyond ammonia and amines in the model. Quantitatively understanding the overall budget, evolution, and thermodynamic properties of lower-volatility organics in the atmosphere will be essential for improving global aerosol models.{\\textless}/p{\\textgreater}},\n\tlanguage = {English},\n\tnumber = {14},\n\turldate = {2018-10-05},\n\tjournal = {Atmospheric Chemistry and Physics},\n\tauthor = {Hodshire, Anna L. and Lawler, Michael J. and Zhao, Jun and Ortega, John and Jen, Coty and Yli-Juuti, Taina and Brewer, Jared F. and Kodros, Jack K. and Barsanti, Kelley C. and Hanson, Dave R. and McMurry, Peter H. and Smith, James N. and Pierce, Jeffery R.},\n\tmonth = jul,\n\tyear = {2016},\n\tpages = {9321--9348},\n}\n\n
\n
\n\n\n
\n \\textlessp\\textgreater\\textlessstrong\\textgreaterAbstract.\\textless/strong\\textgreater New-particle formation (NPF) is a significant source of aerosol particles into the atmosphere. However, these particles are initially too small to have climatic importance and must grow, primarily through net uptake of low-volatility species, from diameters ∼ \\textlessspan class=\"thinspace\"\\textgreater\\textless/span\\textgreater1 to 30–100\\textlessspan class=\"thinspace\"\\textgreater\\textless/span\\textgreaternm in order to potentially impact climate. There are currently uncertainties in the physical and chemical processes associated with the growth of these freshly formed particles that lead to uncertainties in aerosol-climate modeling. Four main pathways for new-particle growth have been identified: condensation of sulfuric-acid vapor (and associated bases when available), condensation of organic vapors, uptake of organic acids through acid–base chemistry in the particle phase, and accretion of organic molecules in the particle phase to create a lower-volatility compound that then contributes to the aerosol mass. The relative importance of each pathway is uncertain and is the focus of this work. \\textlessbr\\textgreater\\textlessbr\\textgreater The 2013 New Particle Formation Study (NPFS) measurement campaign took place at the DOE Southern Great Plains (SGP) facility in Lamont, Oklahoma, during spring 2013. Measured gas- and particle-phase compositions during these new-particle growth events suggest three distinct growth pathways: (1) growth by primarily organics, (2) growth by primarily sulfuric acid and ammonia, and (3) growth by primarily sulfuric acid and associated bases and organics. To supplement the measurements, we used the particle growth model MABNAG (Model for Acid–Base chemistry in NAnoparticle Growth) to gain further insight into the growth processes on these 3 days at SGP. MABNAG simulates growth from (1) sulfuric-acid condensation (and subsequent salt formation with ammonia or amines), (2) near-irreversible condensation from nonreactive extremely low-volatility organic compounds (ELVOCs), and (3) organic-acid condensation and subsequent salt formation with ammonia or amines. MABNAG is able to corroborate the observed differing growth pathways, while also predicting that ELVOCs contribute more to growth than organic salt formation. However, most MABNAG model simulations tend to underpredict the observed growth rates between 10 and 20\\textlessspan class=\"thinspace\"\\textgreater\\textless/span\\textgreaternm in diameter; this underprediction may come from neglecting the contributions to growth from semi-to-low-volatility species or accretion reactions. Our results suggest that in addition to sulfuric acid, ELVOCs are also very important for growth in this rural setting. We discuss the limitations of our study that arise from not accounting for semi- and low-volatility organics, as well as nitrogen-containing species beyond ammonia and amines in the model. Quantitatively understanding the overall budget, evolution, and thermodynamic properties of lower-volatility organics in the atmosphere will be essential for improving global aerosol models.\\textless/p\\textgreater\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Chemical ionization of clusters formed from sulfuric acid and dimethylamine or diamines.\n \n \n \n \n\n\n \n Jen, C. N.; Zhao, J.; McMurry, P. H.; and Hanson, D. R.\n\n\n \n\n\n\n Atmos. Chem. Phys., 16(19): 12513–12529. October 2016.\n \n\n\n\n
\n\n\n\n \n \n \"ChemicalPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{jen_chemical_2016,\n\ttitle = {Chemical ionization of clusters formed from sulfuric acid and dimethylamine or diamines},\n\tvolume = {16},\n\tissn = {1680-7324},\n\turl = {https://www.atmos-chem-phys.net/16/12513/2016/},\n\tdoi = {10.5194/acp-16-12513-2016},\n\tabstract = {Chemical ionization (CI) mass spectrometers are used to study atmospheric nucleation by detecting clusters produced by reactions of sulfuric acid and various basic gases. These instruments typically use nitrate to deprotonate and thus chemically ionize the clusters. In this study, we compare cluster concentrations measured using either nitrate or acetate. Clusters were formed in a flow reactor from vapors of sulfuric acid and dimethylamine, ethylene diamine, tetramethylethylene diamine, or butanediamine (also known as putrescine). These comparisons show that nitrate is unable to chemically ionize clusters with high base content. In addition, we vary the ion–molecule reaction time to probe ion processes which include proton-transfer, ion–molecule clustering, and decomposition of ions. Ion decomposition upon deprotonation by acetate/nitrate was observed. More studies are needed to quantify to what extent ion decomposition affects observed cluster content and concentrations, especially those chemically ionized with acetate since it deprotonates more types of clusters than nitrate.Model calculations of the neutral and ion cluster formation pathways are also presented to better identify the cluster types that are not efficiently deprotonated by nitrate. Comparison of model and measured clusters indicate that sulfuric acid dimers with two diamines and sulfuric acid trimers with two or more base molecules are not efficiently chemical ionized by nitrate. We conclude that acetate CI provides better information on cluster abundancies and their base content than nitrate CI.},\n\tnumber = {19},\n\turldate = {2018-03-26},\n\tjournal = {Atmos. Chem. Phys.},\n\tauthor = {Jen, C. N. and Zhao, J. and McMurry, P. H. and Hanson, D. R.},\n\tmonth = oct,\n\tyear = {2016},\n\tpages = {12513--12529},\n}\n\n
\n
\n\n\n
\n Chemical ionization (CI) mass spectrometers are used to study atmospheric nucleation by detecting clusters produced by reactions of sulfuric acid and various basic gases. These instruments typically use nitrate to deprotonate and thus chemically ionize the clusters. In this study, we compare cluster concentrations measured using either nitrate or acetate. Clusters were formed in a flow reactor from vapors of sulfuric acid and dimethylamine, ethylene diamine, tetramethylethylene diamine, or butanediamine (also known as putrescine). These comparisons show that nitrate is unable to chemically ionize clusters with high base content. In addition, we vary the ion–molecule reaction time to probe ion processes which include proton-transfer, ion–molecule clustering, and decomposition of ions. Ion decomposition upon deprotonation by acetate/nitrate was observed. More studies are needed to quantify to what extent ion decomposition affects observed cluster content and concentrations, especially those chemically ionized with acetate since it deprotonates more types of clusters than nitrate.Model calculations of the neutral and ion cluster formation pathways are also presented to better identify the cluster types that are not efficiently deprotonated by nitrate. Comparison of model and measured clusters indicate that sulfuric acid dimers with two diamines and sulfuric acid trimers with two or more base molecules are not efficiently chemical ionized by nitrate. We conclude that acetate CI provides better information on cluster abundancies and their base content than nitrate CI.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Strong Hydrogen Bonded Molecular Interactions between Atmospheric Diamines and Sulfuric Acid.\n \n \n \n\n\n \n Elm, J.; Jen, C. N.; Kurtén, T.; and Vehkamäki, H.\n\n\n \n\n\n\n The Journal of Physical Chemistry A, 120: 3693–3700. May 2016.\n \n\n\n\n
\n\n\n\n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{elm_strong_2016,\n\ttitle = {Strong {Hydrogen} {Bonded} {Molecular} {Interactions} between {Atmospheric} {Diamines} and {Sulfuric} {Acid}},\n\tvolume = {120},\n\tcopyright = {All rights reserved},\n\tissn = {1089-5639},\n\tdoi = {10.1021/acs.jpca.6b03192},\n\tjournal = {The Journal of Physical Chemistry A},\n\tauthor = {Elm, Jonas and Jen, Coty N. and Kurtén, Theo and Vehkamäki, Hanna},\n\tmonth = may,\n\tyear = {2016},\n\tpages = {3693--3700},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Diamine-sulfuric acid reactions are a potent source of new particle formation.\n \n \n \n\n\n \n Jen, C. N.; Bachman, R.; Zhao, J.; McMurry, P. H.; and Hanson, D. R.\n\n\n \n\n\n\n Geophysical Research Letters,2015GL066958. 2016.\n \n\n\n\n
\n\n\n\n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{jen_diamine-sulfuric_2016,\n\ttitle = {Diamine-sulfuric acid reactions are a potent source of new particle formation},\n\tcopyright = {All rights reserved},\n\tissn = {1944-8007},\n\tdoi = {10.1002/2015GL066958},\n\tabstract = {Atmospheric nucleation from sulfuric acid depends on the concentrations and the stabilizing effect of other trace gases, such as ammonia and amines. Diamines are an understudied class of atmospherically relevant compounds and we examine how they affect sulfuric acid nucleation in both flow reactor experiments and the atmosphere. The number of particles produced from sulfuric acid and diamines in the flow reactor was equal to or greater than the number formed from monoamines, implying that diamines are more effective nucleating agents. Upper limits of diamine abundance were also monitored during three field campaigns: Lamont, OK (2013); Lewes, DE (2012); and Atlanta, GA (2009). Mixing ratios were measured as high as tens of pptv (GA and OK). Laboratory results suggest diamines at these levels are important for atmospheric nucleation. Diamines likely participate in atmospheric nucleation and should be considered in nucleation measurements and models.},\n\tjournal = {Geophysical Research Letters},\n\tauthor = {Jen, Coty N. and Bachman, Ryan and Zhao, Jun and McMurry, Peter H. and Hanson, David R.},\n\tyear = {2016},\n\tkeywords = {0305 Aerosols and particles, amines, atmospheric nucleation, diamines, flow reactor, sulfuric acid},\n\tpages = {2015GL066958},\n}\n\n
\n
\n\n\n
\n Atmospheric nucleation from sulfuric acid depends on the concentrations and the stabilizing effect of other trace gases, such as ammonia and amines. Diamines are an understudied class of atmospherically relevant compounds and we examine how they affect sulfuric acid nucleation in both flow reactor experiments and the atmosphere. The number of particles produced from sulfuric acid and diamines in the flow reactor was equal to or greater than the number formed from monoamines, implying that diamines are more effective nucleating agents. Upper limits of diamine abundance were also monitored during three field campaigns: Lamont, OK (2013); Lewes, DE (2012); and Atlanta, GA (2009). Mixing ratios were measured as high as tens of pptv (GA and OK). Laboratory results suggest diamines at these levels are important for atmospheric nucleation. Diamines likely participate in atmospheric nucleation and should be considered in nucleation measurements and models.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2015\n \n \n (2)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n Sulfuric Acid Nucleation: An Experimental Study of the Effect of Seven Bases.\n \n \n \n\n\n \n Glasoe, W. A.; Volz, K.; Panta, B.; Freshour, N.; Bachman, R.; Hanson, D. R.; McMurry, P. H.; and Jen, C.\n\n\n \n\n\n\n Journal of Geophysical Research: Atmospheres,2014JD022730. 2015.\n \n\n\n\n
\n\n\n\n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n  \n \n 1 download\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{glasoe_sulfuric_2015,\n\ttitle = {Sulfuric {Acid} {Nucleation}: {An} {Experimental} {Study} of the {Effect} of {Seven} {Bases}},\n\tcopyright = {All rights reserved},\n\tissn = {2169-8996},\n\tdoi = {10.1002/2014JD022730},\n\tabstract = {Nucleation of particles with sulfuric acid, water, and nitrogeneous bases was studied in a flow reactor. Sulfuric acid and water levels were set by flows over sulfuric acid and water reservoirs, respectively, and the base concentrations were determined from measured permeation rates and flow dilution ratios. Particle number distributions were measured with a nano- differential mobility analyzer system. Results indicate that the nucleation capability of NH3, methyl-, dimethyl-, and trimethyl- amines with sulfuric acid increases from NH3 as the weakest, methyl amine next, and dimethyl amine and trimethyl amine the strongest. Three other bases were studied and experiments with triethyl amine showed that it is less effective than methyl amine, and experiments with urea and acetamide showed that their capabilities are much lower than the amines with acetamide having basically no effect. When both NH3 and an amine were present, nucleation was more strongly enhanced than with just the amine present. Comparisons of nucleation rates to predictions and previous experimental work are discussed and the sulfuric acid - base nucleation rates measured here are extrapolated to atmospheric conditions. The measurements suggest that atmospheric nucleation rates are significantly affected by synergistic interactions between ammonia and amines.},\n\tjournal = {Journal of Geophysical Research: Atmospheres},\n\tauthor = {Glasoe, W. A. and Volz, K. and Panta, B. and Freshour, N. and Bachman, R. and Hanson, D. R. and McMurry, P. H. and Jen, C.},\n\tyear = {2015},\n\tkeywords = {0305 Aerosols and particles, amines, ammonia, nucleation, sulfuric acid},\n\tpages = {2014JD022730},\n}\n\n
\n
\n\n\n
\n Nucleation of particles with sulfuric acid, water, and nitrogeneous bases was studied in a flow reactor. Sulfuric acid and water levels were set by flows over sulfuric acid and water reservoirs, respectively, and the base concentrations were determined from measured permeation rates and flow dilution ratios. Particle number distributions were measured with a nano- differential mobility analyzer system. Results indicate that the nucleation capability of NH3, methyl-, dimethyl-, and trimethyl- amines with sulfuric acid increases from NH3 as the weakest, methyl amine next, and dimethyl amine and trimethyl amine the strongest. Three other bases were studied and experiments with triethyl amine showed that it is less effective than methyl amine, and experiments with urea and acetamide showed that their capabilities are much lower than the amines with acetamide having basically no effect. When both NH3 and an amine were present, nucleation was more strongly enhanced than with just the amine present. Comparisons of nucleation rates to predictions and previous experimental work are discussed and the sulfuric acid - base nucleation rates measured here are extrapolated to atmospheric conditions. The measurements suggest that atmospheric nucleation rates are significantly affected by synergistic interactions between ammonia and amines.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Towards Reconciling Measurements of Atmospherically Relevant Clusters by Chemical Ionization Mass Spectrometry and Mobility Classification/Vapor Condensation.\n \n \n \n\n\n \n Jen, C. N.; Hanson, D. R.; and McMurry, P. H.\n\n\n \n\n\n\n Aerosol Science and Technology, ARL, 49: i–iii. January 2015.\n \n\n\n\n
\n\n\n\n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{jen_towards_2015,\n\ttitle = {Towards {Reconciling} {Measurements} of {Atmospherically} {Relevant} {Clusters} by {Chemical} {Ionization} {Mass} {Spectrometry} and {Mobility} {Classification}/{Vapor} {Condensation}},\n\tvolume = {49},\n\tcopyright = {All rights reserved},\n\tissn = {0278-6826},\n\tdoi = {10.1080/02786826.2014.1002602},\n\tjournal = {Aerosol Science and Technology, ARL},\n\tauthor = {Jen, Coty N. and Hanson, David R. and McMurry, Peter H.},\n\tmonth = jan,\n\tyear = {2015},\n\tpages = {i--iii},\n}\n\n
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2014\n \n \n (1)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n Stabilization of sulfuric acid dimers by ammonia, methylamine, dimethylamine, and trimethylamine.\n \n \n \n\n\n \n Jen, C. N.; McMurry, P. H.; and Hanson, D. R.\n\n\n \n\n\n\n Journal of Geophysical Research: Atmospheres, 119: 2014JD021592. 2014.\n \n\n\n\n
\n\n\n\n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{jen_stabilization_2014,\n\ttitle = {Stabilization of sulfuric acid dimers by ammonia, methylamine, dimethylamine, and trimethylamine},\n\tvolume = {119},\n\tcopyright = {All rights reserved},\n\tissn = {2169-8996},\n\tdoi = {10.1002/2014JD021592},\n\tjournal = {Journal of Geophysical Research: Atmospheres},\n\tauthor = {Jen, Coty N. and McMurry, Peter H. and Hanson, David R.},\n\tyear = {2014},\n\tkeywords = {0305 Aerosols and particles, 0317 Chemical kinetic and photochemical properties, 0394 Instruments and techniques, 3305 Climate change and variability, 3311 Clouds and aerosols, aerosol particles, amines, atmospheric nucleation, dimer, flow reactor, mass spectrometry},\n\tpages = {2014JD021592},\n}\n\n
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2013\n \n \n (1)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n Quantitative and time-resolved nanoparticle composition measurements during new particle formation.\n \n \n \n\n\n \n Bzdek, B. R.; Horan, A. J.; Pennington, M. R.; DePalma, J. W.; Zhao, J.; Jen, C. N.; Hanson, D. R.; Smith, J. N.; McMurry, P. H.; and Johnston, M. V.\n\n\n \n\n\n\n Faraday Discussions, 165: 25–43. 2013.\n \n\n\n\n
\n\n\n\n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{bzdek_quantitative_2013,\n\ttitle = {Quantitative and time-resolved nanoparticle composition measurements during new particle formation},\n\tvolume = {165},\n\tcopyright = {All rights reserved},\n\tissn = {1359-6640},\n\tdoi = {10.1039/C3FD00039G},\n\tabstract = {The chemical composition of 20 nm diameter particles was measured with the Nano Aerosol Mass Spectrometer (NAMS) in a rural/coastal environment during days when new particle formation (NPF) occurred and days when NPF did not occur. NAMS provides a quantitative measure of nanoparticle elemental composition with high time resolution. These measurements show that nanoparticle chemical composition is dynamic on both types of days and that changes in nanoparticle chemical composition do not necessarily correlate with changes in aerosol mass or number concentration. On NPF days, NAMS can distinguish between elements associated with particle formation and early mass growth from those associated with later mass growth. In the early stage of NPF, the particle phase sulphur mole fraction (S) increases simultaneously with the increase in gas phase sulphuric acid. This composition change occurs before the mode diameter has grown into the NAMS-measured size range and is quantitatively described by sulphuric acid condensation. The nitrogen mole fraction (N) also increases during this time period. The N/S mole ratio is approximately 2, indicating that particulate sulphate is fully neutralized. As the mode diameter passes into and through the NAMS-measured size range, N increases at a faster rate than S (N/S mole ratio increases above 2), indicating that a separate, nitrogen-based growth process exists, possibly involving aminium salts, inorganic nitrate and/or organonitrates. Carbonaceous matter is the most abundant component ([similar]50\\% by mass) of the growing nanoparticles, but it is the inorganic species that are preferentially enhanced during NPF relative to other times of day. Concurrent measurements of cloud condensation nucleation activity during NPF events suggest that these newly formed particles are hygroscopic. Nanoparticle composition on non-NPF days also shifts toward a more inorganic composition during the daytime, but the chemical species are different from NPF days and the particles are less hygroscopic. Incorporation of S into growing nanoparticles is adequately explained by existing models, but currently no models exist to satisfactorily explain incorporation of nitrogen-containing species or carbonaceous matter.},\n\tjournal = {Faraday Discussions},\n\tauthor = {Bzdek, Bryan R. and Horan, Andrew J. and Pennington, M. Ross and DePalma, Joseph W. and Zhao, Jun and Jen, Coty N. and Hanson, David R. and Smith, James N. and McMurry, Peter H. and Johnston, Murray V.},\n\tyear = {2013},\n\tpages = {25--43},\n}\n
\n
\n\n\n
\n The chemical composition of 20 nm diameter particles was measured with the Nano Aerosol Mass Spectrometer (NAMS) in a rural/coastal environment during days when new particle formation (NPF) occurred and days when NPF did not occur. NAMS provides a quantitative measure of nanoparticle elemental composition with high time resolution. These measurements show that nanoparticle chemical composition is dynamic on both types of days and that changes in nanoparticle chemical composition do not necessarily correlate with changes in aerosol mass or number concentration. On NPF days, NAMS can distinguish between elements associated with particle formation and early mass growth from those associated with later mass growth. In the early stage of NPF, the particle phase sulphur mole fraction (S) increases simultaneously with the increase in gas phase sulphuric acid. This composition change occurs before the mode diameter has grown into the NAMS-measured size range and is quantitatively described by sulphuric acid condensation. The nitrogen mole fraction (N) also increases during this time period. The N/S mole ratio is approximately 2, indicating that particulate sulphate is fully neutralized. As the mode diameter passes into and through the NAMS-measured size range, N increases at a faster rate than S (N/S mole ratio increases above 2), indicating that a separate, nitrogen-based growth process exists, possibly involving aminium salts, inorganic nitrate and/or organonitrates. Carbonaceous matter is the most abundant component ([similar]50% by mass) of the growing nanoparticles, but it is the inorganic species that are preferentially enhanced during NPF relative to other times of day. Concurrent measurements of cloud condensation nucleation activity during NPF events suggest that these newly formed particles are hygroscopic. Nanoparticle composition on non-NPF days also shifts toward a more inorganic composition during the daytime, but the chemical species are different from NPF days and the particles are less hygroscopic. Incorporation of S into growing nanoparticles is adequately explained by existing models, but currently no models exist to satisfactorily explain incorporation of nitrogen-containing species or carbonaceous matter.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2012\n \n \n (1)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n Acid–base chemical reaction model for nucleation rates in the polluted atmospheric boundary layer.\n \n \n \n\n\n \n Chen, M.; Titcombe, M.; Jiang, J.; Jen, C.; Kuang, C.; Fischer, M. L.; Eisele, F. L.; Siepmann, J. I.; Hanson, D. R.; Zhao, J.; and McMurry, P. H.\n\n\n \n\n\n\n Proceedings of the National Academy of Sciences, 109: 18713–18718. October 2012.\n \n\n\n\n
\n\n\n\n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{chen_acidbase_2012,\n\ttitle = {Acid–base chemical reaction model for nucleation rates in the polluted atmospheric boundary layer},\n\tvolume = {109},\n\tcopyright = {All rights reserved},\n\tdoi = {10.1073/pnas.1210285109},\n\tabstract = {Climate models show that particles formed by nucleation can affect cloud cover and, therefore, the earth's radiation budget. Measurements worldwide show that nucleation rates in the atmospheric boundary layer are positively correlated with concentrations of sulfuric acid vapor. However, current nucleation theories do not correctly predict either the observed nucleation rates or their functional dependence on sulfuric acid concentrations. This paper develops an alternative approach for modeling nucleation rates, based on a sequence of acid–base reactions. The model uses empirical estimates of sulfuric acid evaporation rates obtained from new measurements of neutral molecular clusters. The model predicts that nucleation rates equal the sulfuric acid vapor collision rate times a prefactor that is less than unity and that depends on the concentrations of basic gaseous compounds and preexisting particles. Predicted nucleation rates and their dependence on sulfuric acid vapor concentrations are in reasonable agreement with measurements from Mexico City and Atlanta.},\n\tjournal = {Proceedings of the National Academy of Sciences},\n\tauthor = {Chen, Modi and Titcombe, Mari and Jiang, Jingkun and Jen, Coty and Kuang, Chongai and Fischer, Marc L. and Eisele, Fred L. and Siepmann, J. Ilja and Hanson, David R. and Zhao, Jun and McMurry, Peter H.},\n\tmonth = oct,\n\tyear = {2012},\n\tkeywords = {amines, atmospheric aerosol, chamber study, climate forcing, nanoparticle},\n\tpages = {18713--18718},\n}\n\n
\n
\n\n\n
\n Climate models show that particles formed by nucleation can affect cloud cover and, therefore, the earth's radiation budget. Measurements worldwide show that nucleation rates in the atmospheric boundary layer are positively correlated with concentrations of sulfuric acid vapor. However, current nucleation theories do not correctly predict either the observed nucleation rates or their functional dependence on sulfuric acid concentrations. This paper develops an alternative approach for modeling nucleation rates, based on a sequence of acid–base reactions. The model uses empirical estimates of sulfuric acid evaporation rates obtained from new measurements of neutral molecular clusters. The model predicts that nucleation rates equal the sulfuric acid vapor collision rate times a prefactor that is less than unity and that depends on the concentrations of basic gaseous compounds and preexisting particles. Predicted nucleation rates and their dependence on sulfuric acid vapor concentrations are in reasonable agreement with measurements from Mexico City and Atlanta.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2009\n \n \n (1)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Light-absorbing secondary organic material formed by glyoxal in aqueous aerosol mimics.\n \n \n \n \n\n\n \n Shapiro, E. L.; Szprengiel, J.; Sareen, N.; Jen, C. N.; Giordano, M. R.; and McNeill, V. F.\n\n\n \n\n\n\n Atmospheric Chemistry and Physics, 9(7): 2289–2300. April 2009.\n \n\n\n\n
\n\n\n\n \n \n \"Light-absorbingPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{shapiro_light-absorbing_2009,\n\ttitle = {Light-absorbing secondary organic material formed by glyoxal in aqueous aerosol mimics},\n\tvolume = {9},\n\tcopyright = {All rights reserved},\n\tissn = {1680-7316},\n\turl = {https://www.atmos-chem-phys.net/9/2289/2009/},\n\tdoi = {https://doi.org/10.5194/acp-9-2289-2009},\n\tabstract = {{\\textless}p{\\textgreater}{\\textless}strong{\\textgreater}Abstract.{\\textless}/strong{\\textgreater} Light-absorbing and high-molecular-weight secondary organic products were observed to result from the reaction of glyoxal in mildly acidic (pH=4) aqueous inorganic salt solutions mimicking aqueous tropospheric aerosol particles. High-molecular-weight (500–600 amu) products were observed when ammonium sulfate ((NH$_{\\textrm{4}}$)$_{\\textrm{2}}$SO$_{\\textrm{4}}$) or sodium chloride (NaCl) was present in the aqueous phase. The products formed in (NH$_{\\textrm{4}}$)$_{\\textrm{2}}$SO$_{\\textrm{4}}$ or ammonium nitrate (NH$_{\\textrm{4}}$NO$_{\\textrm{3}}$) solutions absorb light at UV and visible wavelengths. Substantial absorption at 300–400 nm develops within two hours, and absorption between 400–600 nm develops within days. Pendant drop tensiometry measurements show that the products are not surface-active. The experimental results along with ab initio predictions of the UV/Vis absorption of potential products suggest a mechanism involving the participation of the ammonium ion. If similar products are formed in atmospheric aerosol particles, they could change the optical properties of the seed aerosol over its lifetime.{\\textless}/p{\\textgreater}},\n\tlanguage = {English},\n\tnumber = {7},\n\turldate = {2019-01-18},\n\tjournal = {Atmospheric Chemistry and Physics},\n\tauthor = {Shapiro, E. L. and Szprengiel, J. and Sareen, N. and Jen, C. N. and Giordano, M. R. and McNeill, V. F.},\n\tmonth = apr,\n\tyear = {2009},\n\tpages = {2289--2300},\n}\n\n
\n
\n\n\n
\n \\textlessp\\textgreater\\textlessstrong\\textgreaterAbstract.\\textless/strong\\textgreater Light-absorbing and high-molecular-weight secondary organic products were observed to result from the reaction of glyoxal in mildly acidic (pH=4) aqueous inorganic salt solutions mimicking aqueous tropospheric aerosol particles. High-molecular-weight (500–600 amu) products were observed when ammonium sulfate ((NH$_{\\textrm{4}}$)$_{\\textrm{2}}$SO$_{\\textrm{4}}$) or sodium chloride (NaCl) was present in the aqueous phase. The products formed in (NH$_{\\textrm{4}}$)$_{\\textrm{2}}$SO$_{\\textrm{4}}$ or ammonium nitrate (NH$_{\\textrm{4}}$NO$_{\\textrm{3}}$) solutions absorb light at UV and visible wavelengths. Substantial absorption at 300–400 nm develops within two hours, and absorption between 400–600 nm develops within days. Pendant drop tensiometry measurements show that the products are not surface-active. The experimental results along with ab initio predictions of the UV/Vis absorption of potential products suggest a mechanism involving the participation of the ammonium ion. If similar products are formed in atmospheric aerosol particles, they could change the optical properties of the seed aerosol over its lifetime.\\textless/p\\textgreater\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n\n\n\n
\n\n\n \n\n \n \n \n \n\n
\n"}; document.write(bibbase_data.data);