Natural Marine Precursors Boost Continental New Particle Formation and Production of Cloud Condensation NucleiClick to copy article linkArticle link copied!
- Robin Wollesen de Jonge*Robin Wollesen de Jonge*Email: [email protected]Department of Physics, Lund University, Professorsgatan 1, Lund SE-22363, SwedenMore by Robin Wollesen de Jonge
- Carlton XavierCarlton XavierDepartment of Physics, Lund University, Professorsgatan 1, Lund SE-22363, SwedenSwedish Meteorological and Hydrological Institute (SMHI), Norrköping SE-60176, SwedenMore by Carlton Xavier
- Tinja OleniusTinja OleniusSwedish Meteorological and Hydrological Institute (SMHI), Norrköping SE-60176, SwedenMore by Tinja Olenius
- Jonas ElmJonas ElmDepartment of Chemistry, Aarhus University, Langelandsgade 140, Aarhus DK-8000, DenmarkMore by Jonas Elm
- Carl SvenhagCarl SvenhagDepartment of Physics, Lund University, Professorsgatan 1, Lund SE-22363, SwedenMore by Carl Svenhag
- Noora HyttinenNoora HyttinenFinnish Meteorological Institute, Kuopio FI-70211, FinlandDepartment of Chemistry, Nanoscience Center, University of Jyväskylä, Jyväskylä FI-40014, FinlandMore by Noora Hyttinen
- Lars NieradzikLars NieradzikDepartment of Physical Geography and Ecosystem Science, Lund University, Lund SE-22362, SwedenMore by Lars Nieradzik
- Nina SarnelaNina SarnelaInstitute for Atmospheric and Earth System Research/Physics, Faculty of Science, University of Helsinki, Helsinki FI-00014, FinlandMore by Nina Sarnela
- Adam KristenssonAdam KristenssonDepartment of Physics, Lund University, Professorsgatan 1, Lund SE-22363, SwedenMore by Adam Kristensson
- Tuukka PetäjäTuukka PetäjäInstitute for Atmospheric and Earth System Research/Physics, Faculty of Science, University of Helsinki, Helsinki FI-00014, FinlandJoint International Research Laboratory of Atmospheric and Earth System Sciences, School of Atmospheric Sciences, Nanjing University, Nanjing CN-210023, ChinaMore by Tuukka Petäjä
- Mikael EhnMikael EhnInstitute for Atmospheric and Earth System Research/Physics, Faculty of Science, University of Helsinki, Helsinki FI-00014, FinlandMore by Mikael Ehn
- Pontus RoldinPontus RoldinDepartment of Physics, Lund University, Professorsgatan 1, Lund SE-22363, SwedenSwedish Environmental Research Institute IVL, Malmö SE-21119, SwedenMore by Pontus Roldin
Abstract
Marine dimethyl sulfide (DMS) emissions are the dominant source of natural sulfur in the atmosphere. DMS oxidizes to produce low-volatility acids that potentially nucleate to form particles that may grow into climatically important cloud condensation nuclei (CCN). In this work, we utilize the chemistry transport model ADCHEM to demonstrate that DMS emissions are likely to contribute to the majority of CCN during the biological active period (May-August) at three different forest stations in the Nordic countries. DMS increases CCN concentrations by forming nucleation and Aitken mode particles over the ocean and land, which eventually grow into the accumulation mode by condensation of low-volatility organic compounds from continental vegetation. Our findings provide a new understanding of the exchange of marine precursors between the ocean and land, highlighting their influence as one of the dominant sources of CCN particles over the boreal forest.
This publication is licensed under
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
Note Added after ASAP Publication
Due to a production error, this paper was published ASAP on June 13, 2024, with an error in Table 3. The corrected version was reposted on June 25, 2024.
Synopsis
Little is known on the role of natural marine precursors in the formation of aerosol particles over land. Here we demonstrate that dimethyl sulfide in particular gives rise to a substantial fraction of aerosol particles in the cloud condensation nuclei size range observed over the Nordic boreal forest.
1. Introduction
2. Materials and Methods
ADCHEM Model Description
Simulations Period, Location, and Measurement Data
Gas and Primary Particle Emissions
New Particle Formation
Henry’s Law Solubility and pKa
Adiabatic Cloud Parcel Model
Air-Mass Origin Analysis
Statistical Methods
3. Results and Discussion
Air Mass History and Aerosol Size Distributions
Figure 1
Figure 1. Effect of marine versus continental air mass history on the aerosol number size distribution in a boreal forest environment. Panel (a) depicts the typical summertime source regions of DMS (purple), MTs (green), NH3 (blue), and anthropogenic SO2 (yellow), along with their position in relation to the Hyltemossa, SMEARII, and Pallas measurements stations. These colors indicate only the spatial locations of the emissions of the different species between May and August, 2018. Panels (b), (c), and (d) show the spring and summertime (March-August) mean aerosol number size distributions measured at the Hyltemossa, SMEARII, and Pallas stations during the periods 2018–2020, 2006–2020, and 2015–2018, respectively. The size distributions in panel (b), (c,) and (d) are separated based on the marine influence on their air mass history.
Marine Air-Mass Influence on the Gas-Phase Chemistry over the Boreal Forest
Figure 2
Figure 2. Modeled and measured long-term and diurnal gas-phase concentrations of MTs, HOM monomers, isoprene, SA, MSA, and O3 at the Station for Measuring Ecosystem-Atmosphere Relations II (SMEARII) between the 17th of May and 28th of August, 2018. Gray and black dots denote the surface and above-canopy measurements, respectively, while the colored lines represent the above-canopy model results. Marine periods 1, 2, and 3 (MP1, MP2, and MP3) denote periods of particularly high marine air-mass impact, while continental period 1 (CP1) denotes a period of particularly high continental air-mass impact. The back-trajectory heat-maps displayed above each period illustrate the regions from which the air-masses arrived during MP1, MP2, MP3, and CP1. The shaded areas denote the measured and modeled data range within the 25th and 75th percentile.
Species | O̅ (cm–3)b | M̅ (cm–3) | R | NMB | FAC2 |
---|---|---|---|---|---|
MT | 3.74 × 109 | 2.20 × 109 | 0.32 (0.81) | –0.48 (−0.38) | 0.45 (1.00) |
HOMm | 1.09 × 108 | 1.37 × 107 | 0.39 (0.23) | –0.87 (−0.87) | 0.03 (0.00) |
Isoprene | 3.31 × 109 | 3.49 × 109 | 0.36 (0.38) | 0.31 (−0.01) | 0.45 (1.0) |
SA | 5.25 × 105 | 1.10 × 106 | 0.24 (0.96) | 2.19 (2.19) | 0.23 (0.25) |
MSA | 1.82 × 105 | 1.80 × 105 | 0.29 (0.84) | 5.69 (2.17) | 0.17 (0.50) |
O3 | 32.5 (ppb) | 30.66 (ppb) | 0.39 (0.61) | –0.03 (−0.18) | 0.86 (1.00) |
A comparison between the modeled and measured concentrations is provided for the full period between May and August and for the diurnal variability in the gas-phase concentrations averaged over the entire period.
O̅denotes the observed median gas-phase concentration of each species,M̅ the modeled median gas-phase concentration, R the correlation coefficient, NMB the normalized mean bias, and FAC2 the fraction of predictions within a factor two of the observations. Diurnal results are reported in parentheses.
Marine Air-Mass Impact on Particle Formation and Growth over the Boreal Forest
Figure 3
Figure 3. Measured and modeled (a–e) time-dependent and (f–o) median particle number size distributions at the Station for Measuring Ecosystem Atmosphere Relations II (SMEARII) between the 17th of May and 28th of August. Marine periods 1, 2, and 3 (MP1, MP2, and MP3) denote periods of particular high marine air-mass impact, while continental period 1 (CP1) denotes a period of particular high continental air-mass impact. The back-trajectory heat-maps displayed above each period illustrates the regions from which the air-masses arrived during MP1, MP2, MP3, and CP1. The median size distributions are separated into (f–j) periods of predominant marine air-mass impact (time spent over the ocean >50th prct., purple), and (k–o) periods of predominant continental impact (time spent over the ocean <50th prct., green). The averaging for the median size distributions is done for the whole simulation period and not just during MP1, MP2, MP3, and CP1. The model results include data from the base case run (BaseCase), the without DMS emissions simulation (woDMS), the without iodine nucleation simulation (woIodine), and the without anthropogenic emissions simulation (woAnthro). The shaded areas denote the measured and modeled data range within the 25th and 75th percentile.
Impact of Marine Species on Clouds and Climate Over the Boreal Forest
Figure 4
Figure 4. Schematic depiction of the influence of DMS and various iodine species on the formation and growth of aerosol particles both over the ocean and over land. (1) DMS, CH3I, I2, and HOI oxidize over the ocean and land to form the strong acids SA, MSA, HIO3, and HIO2, which nucleate (3) among themselves or (2) in the presence of NH3 or DMA. These particles in turn are (4) grown by condensation of low volatile organic compounds over the boreal forest, ultimately reaching (5) the CCN size range, where they (6) affect the formation, lifetime, and precipitation of clouds. Created with BioRender.com.
CCN change at w = 0.1 ms–1 (%) | CCN change at w = 1.0 ms–1 (%) | |||||||
---|---|---|---|---|---|---|---|---|
Model Run | MP1 | MP2 | MP3 | CP1 | MP1 | MP2 | MP3 | CP1 |
woDMS | –78.3 | –81.5 | –68.6 | –6.0% | –60.8 | –39.4 | –51.2 | –2.8 |
woIodinea | 4.0 | –2.4 | 4.2 | –3.9 | –2.1 | –2.5 | –0.3 | –8.4 |
woAnthro | 8.0 | 5.6 | 26.8 | –14.2 | –32.6 | –30.3 | –31.7 | –23.3 |
Iodine nucleation includes HIO3-HIO2 and HIO3-DMA. Anthroprogenic emissions comprise SO2, NOx, NH3, CO, BC, and AVOCs.
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.est.4c01891.
Figure S1 contains modeled and measured gas-phase concentrations of SO2, HIO3, and NOx, along with modeled gas-phase concentrations of OH, NH3, and DMA; Figure S2 illustrates the back trajectory heat maps for the full simulation period along with MP1, MP2, MP3, and CP1; Figure S3 demonstrates the method behind the air-mass origin analysis; Figure S4 contains the modeled size resolved chemical composition of SO4, NO3, Cl, NH4, Na, MSA, HIO3, DMA, SOA, and POA between 1 nm and 1 μm; Figure S5 illustrates the modeled and measured median particle number size distribution, including results from the woDMS, woIodine, and woAnthro sensitivity runs; Figure S6 contains modeled and measured particle numbers size distributions from Pallas, while Figure S7 contains similar results from Hyltemossa; Figure S8 illustrates the DMS, SO2, and OH gas-phase concentration along with the SA-NH3 nucleation rate along one of the HYSPLIT trajectories moving from the Norwegian Sea toward the SMEARII stationl Table S1 comprises COSMOtherm-derived Henry’s law solubilities, while Table S2 contains COSMOtherm-derived pKa values (PDF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.
Acknowledgments
The authors would like to thank Markku Kulmala and Pasi Aalto for DMPS measurements obtained at SMEARII and Niku Kivekäs and Sami Seppälä for DMPS measurements obtained at Pallas. This project has received funding from the Swedish Research Council Formas (project no. 2018-01745-COBACCA), the Swedish Research Council VR (project no. 2019-05006), the Crafoord foundation (project no. 20210969), the Horizon Europe project AVENGERS (project no. 101081322) and the Academy of Finland (grant no. 338171). We thank the Swedish Strategic Research Program MERGE, the Profile Area Aerosols at the Faculty of Engineering at Lund University and the Profile Area Nature-Based Future Solutions at Lund University for strategic support. We gratefully acknowledge the Centre for Scientific and Technical Computing at Lund University, LUNARC, the Swedish National Infrastructure for Computing, SNIC and CSC - IT Center for Science, Finland, for computational resources. We thank ECCAD for archiving and distribution of data from CAMS. LUNARC is partially funded by the Swedish Research Council through grant agreement no. 2016-07213. J.E. thanks the Independent Research Fund Denmark grant number 9064-00001B for financial support. T.O. acknowledges the Swedish Research Council VR (grant no. 2019-04853) and the Swedish Research Council for Sustainable Development FORMAS (grant no. 2019-01433) for financial support. Funding through the European Commission Horizon Europe project FOCI,20 ”Non-CO2 Forcers and Their Climate, Weather, Air Quality and Health Impacts (project 101056783), FORCeS (grant agreement 821205) and Academy of Finland (ACCC Flagship, project 337549; academy projects 334792, 325681, 333397. The observations at SMEAR II are supported via Academy of Finland (328616, 345510) and via University of Helsinki (HY-ACTRIS).
References
This article references 72 other publications.
- 1Lovelock, J. E.; Maggs, R. J.; Rasmussen, R. A. Atmospheric Dimethyl Sulfide and the Natural Sulfur Cycle. Nature 1972, 237, 452– 453, DOI: 10.1038/237452a0Google Scholar1https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE3sXisVOmug%253D%253D&md5=c6d0495b981f24e9fef6102dd86617eeAtmospheric dimethyl sulfide and the natural sulfur cycleLovelock, J. E.; Maggs, R. J.; Rasmussen, R. A.Nature (London, United Kingdom) (1972), 237 (5356), 452-3CODEN: NATUAS; ISSN:0028-0836.The av. concn. of dimethyl sulfide (DMS) [75-18-3] in seawater was 1.2 .tim. 10-11 g/ml and the distribution coeff. of DMS between air and seawater was 0.30, corresponding to an atm. concn. of 1.2 ppb at equil. The emission rates of DMS were 50-400 .tim. 10-12 g/g wet wt./hr from marine algae, 21-84 .tim. 10-12 g/g/hr from soils, and 2-43 .tim. 10-12 g/g dry wt./hr from the living intact leaves of oak, cotton, spruce, and pine trees. DMS may be involved in the transfer of biol. sulfur [7704-34-9] in nature.
- 2Carpenter, L.; Archer, S.; Beale, R. Ocean-atmosphere trace gas exchange. Chem. Soc. Rev. 2012, 41, 6473– 506, DOI: 10.1039/c2cs35121hGoogle Scholar2https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xhtlaktr%252FP&md5=12993bba4892e6ceadc7b6dc4d8e9432Ocean-atmosphere trace gas exchangeCarpenter, Lucy J.; Archer, Stephen D.; Beale, RachaelChemical Society Reviews (2012), 41 (19), 6473-6506CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. The oceans contribute significantly to the global emissions of a no. of atmospherically important volatile gases, notably those contg. sulfur, nitrogen, and halogens. Such gases play crit. roles not only in global biogeochem. cycling but also in a wide range of atm. processes including marine aerosol formation and modification, tropospheric ozone formation and destruction, photooxidant cycling, and stratospheric ozone loss. A no. of marine emissions are greenhouse gases, others influence the Earth's radiative budget indirectly through aerosol formation and/or by modifying oxidant levels and thus changing the atm. lifetime of gases such as methane. Reviewed is current literature concerning the phys., chem., and biol. controls on the sea-air emissions of a wide range of gases including di-Me sulfide (DMS), halocarbons, nitrogen-contg. gases including NH3, amines (including dimethylamine, DMA, and diethylamine, DEA), alkyl nitrates (RONO2) and N2O, non-methane hydrocarbons (NMHC) including isoprene and oxygenated (O)VOCs, CH4 and CO. Where possible the current global emission budgets of these gases as well as known mechanisms for their formation and loss in the surface ocean are reviewed.
- 3Zheng, G.; Wang, Y.; Wood, R.; Jensen, M. P.; Kuang, C.; McCoy, I. L.; Matthews, A.; Mei, F.; Tomlinson, J. M.; Shilling, J. E. New particle formation in the remote marine boundary layer. Nat. Commun. 2021, 12, 527, DOI: 10.1038/s41467-020-20773-1Google Scholar3https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXit1alsb4%253D&md5=cb0000d7b4413432690fd0ddf779d694New particle formation in the remote marine boundary layerZheng, Guangjie; Wang, Yang; Wood, Robert; Jensen, Michael P.; Kuang, Chongai; McCoy, Isabel L.; Matthews, Alyssa; Mei, Fan; Tomlinson, Jason M.; Shilling, John E.; Zawadowicz, Maria A.; Crosbie, Ewan; Moore, Richard; Ziemba, Luke; Andreae, Meinrat O.; Wang, JianNature Communications (2021), 12 (1), 527CODEN: NCAOBW; ISSN:2041-1723. (Nature Research)Abstr.: Marine low clouds play an important role in the climate system, and their properties are sensitive to cloud condensation nuclei concns. While new particle formation represents a major source of cloud condensation nuclei globally, the prevailing view is that new particle formation rarely occurs in remote marine boundary layer over open oceans. Here we present evidence of the regular and frequent occurrence of new particle formation in the upper part of remote marine boundary layer following cold front passages. The new particle formation is facilitated by a combination of efficient removal of existing particles by pptn., cold air temps., vertical transport of reactive gases from the ocean surface, and high actinic fluxes in a broken cloud field. The newly formed particles subsequently grow and contribute substantially to cloud condensation nuclei in the remote marine boundary layer and thereby impact marine low clouds.
- 4Barnes, I.; Hjorth, J.; Mihalopoulos, N. Dimethyl Sulfide and Dimethyl Sulfoxide and Their Oxidation in the Atmosphere. Chem. Rev. 2006, 106, 940– 975, DOI: 10.1021/cr020529+Google Scholar4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28Xhs1Grtr4%253D&md5=c2a51fd2de3a72676c30eb83aeed4fb1Dimethyl Sulfide and Dimethyl Sulfoxide and Their Oxidation in the AtmosphereBarnes, Ian; Hjorth, Jens; Mihalopoulos, NikosChemical Reviews (Washington, DC, United States) (2006), 106 (3), 940-975CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review concerning the gas phase chem., atm. oxidn., and chem. products of the atm. oxidn. of di-Me sulfide (DMS) and dimethylsulfoxide (DMSO) is given. Topics discussed include: DMS chem. (reaction with OH-, NO3-, halogen atoms and halogen oxides [kinetics, primary reaction mechanisms, reaction products]); DMSO chem. (reaction with OH-, NO3-, halogen atoms and halogen oxides [kinetics, primary reaction mechanisms, reaction products]); and current state of field measurements and their interpretation (multi-phase chem. involved in atm. oxidn. of DMS; field and modeling study evidence of the role of multi-phase reactions in the DMS cycle; aq.-phase reactions of DMS, DMSO, di-Me sulfone, methane sulfinic acid [MSIA], and methane sulfonic acid [MSA]; atm. implication of aq.-phase reactions of DMS, DMSO, MSIA, and MSA; and modeling study recommendations).
- 5Hoffmann, E. H.; Tilgner, A.; Schrödner, R.; Bräuer, P.; Wolke, R.; Herrmann, H. An advanced modeling study on the impacts and atmospheric implications of multiphase dimethyl sulfide chemistry. Proc. Natl. Acad. Sci. U. S. A. 2016, 113, 11776– 11781, DOI: 10.1073/pnas.1606320113Google Scholar5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XhsFyrsb%252FO&md5=97c54638e0a27847d0f959788895cd49An advanced modeling study on the impacts and atmospheric implications of multiphase dimethyl sulfide chemistryHoffmann, Erik Hans; Tilgner, Andreas; Schroedner, Roland; Braeuer, Peter; Wolke, Ralf; Herrmann, HartmutProceedings of the National Academy of Sciences of the United States of America (2016), 113 (42), 11776-11781CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)Oceans dominate di-Me sulfide (DMS) emissions, the major natural S source. DMS is important for formation of non-sea salt sulfate (nss-SO42-) aerosols and secondary particulate matter over oceans; thus, it significantly affects global climate. The DMS oxidn. mechanism has been examd. in several different model studies; however, these studies had restricted oxidn. mechanisms which mainly under-represented important aq.-phase chem. processes. These neglected but highly effective processes strongly affect the direct product yields of DMS oxidn., thereby affecting the climatic influence of aerosols. To address these shortfalls, an extensive multiphase DMS chem. mechanism, the Chem. Aq. Phase Radical Mechanism DMS Module 1.0, was developed and used in detailed model assessments of multiphase DMS chem. in the marine boundary layer. Model studies confirmed the importance of aq. phase chem. for the fate of DMS and its oxidn. products. Aq. phase processes significantly reduced the SO2 yield and increased the Me sulfonic acid (MSA) yield, which is needed to close the gap between modeled and measured MSA concns. Simulations implied that multi-phase DMS oxidn. produces equal amts. of MSA and SO42-, a result with significant implications for nss-SO42- aerosol formation, cloud condensation nuclei concn., and cloud albedo over oceans. Results showed the deficiencies of parameterizations currently used in higher-scale models which only treat gas phase chem. Overall, the results showed treatment of DMS chem. in gas and aq. phases is essential to improve model prediction accuracy.
- 6Wollesen de Jonge, R.; Elm, J.; Rosati, B.; Christiansen, S.; Hyttinen, N.; Lüdemann, D.; Bilde, M.; Roldin, P. Secondary aerosol formation from dimethyl sulfide - improved mechanistic understanding based on smog chamber experiments and modelling. Atmospheric Chemistry and Physics 2021, 21, 9955– 9976, DOI: 10.5194/acp-21-9955-2021Google ScholarThere is no corresponding record for this reference.
- 7Gomez Martin, J. C.; Lewis, T. R.; Blitz, M. A.; Plane, J. M. C.; Kumar, M.; Francisco, J. S.; Saiz-Lopez, A. A gas-to-particle conversion mechanism helps to explain atmospheric particle formation through clustering of iodine oxides. Nat. Commun. 2020, 11, 4521, DOI: 10.1038/s41467-020-18252-8Google ScholarThere is no corresponding record for this reference.
- 8Finkenzeller, H.; Iyer, S.; He, X.-C.; Simon, M.; Koenig, T. K.; Lee, C. F.; Valiev, R.; Hofbauer, V.; Amorim, A.; Baalbaki, R. The gas-phase formation mechanism of iodic acid as an atmospheric aerosol source. Nat. Chem. 2023, 15, 129– 135, DOI: 10.1038/s41557-022-01067-zGoogle ScholarThere is no corresponding record for this reference.
- 9Korhonen, P.; Kulmala, M.; Laaksonen, A.; Viisanen, Y.; McGraw, R.; Seinfeld, J. H. Ternary nucleation of H2SO4, NH3, and H2O in the atmosphere. Journal of Geophysical Research: Atmospheres 1999, 104, 26349– 26353, DOI: 10.1029/1999JD900784Google Scholar9https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXotFWqurk%253D&md5=d5ca8b589881057516ae0bb3ec62b4a9Ternary nucleation of H2SO4, NH3, and H2O in the atmosphereKorhonen, P.; Kulmala, M.; Laaksonen, A.; Viisanen, Y.; McGraw, R.; Seinfeld, J. H.Journal of Geophysical Research, [Atmospheres] (1999), 104 (D21), 26349-26353CODEN: JGRDE3 ISSN:. (American Geophysical Union)Classical theory of binary homogeneous nucleation is extended to the ternary system H2SO4-NH3-H2O. For NH3 mixing ratios exceeding about 1 ppt, the presence of NH3 enhances the binary H2SO4-H2O nucleation rate by several orders of magnitude. The Gibbs free energies of formation of the crit. H2SO4-NH3-H2O cluster, as calcd. by two independent approaches, are in substantial agreement. The finding that the H2SO4-NH3-H2O ternary nucleation rate is independent of relative humidity over a large range of H2SO4 concns. has wide atm. consequences. The limiting component for ternary H2SO4-NH3-H2O nucleation is, as in the binary H2SO4-H2O case, H2SO4; however, the H2SO4 concn. needed to achieve significant nucleation rates is several orders of magnitude below that required in the binary case.
- 10Almeida, J.; Schobesberger, S.; Kürten, A.; Ortega, I.; Kupiainen-Määttä, O.; Praplan, A.; Adamov, A.; Amorim, A.; Bianchi, F.; Breitenlechner, M. Molecular understanding of sulphuric acid-amine particle nucleation in the atmosphere. Nature 2013, 502, 359– 363, DOI: 10.1038/nature12663Google Scholar10https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhsFOit73N&md5=06460967339b7bcd3d231102fd2d96dbMolecular understanding of sulphuric acid-amine particle nucleation in the atmosphereAlmeida, Joao; Schobesberger, Siegfried; Kuerten, Andreas; Ortega, Ismael K.; Kupiainen-Maeaettae, Oona; Praplan, Arnaud P.; Adamov, Alexey; Amorim, Antonio; Bianchi, Federico; Breitenlechner, Martin; David, Andre; Dommen, Josef; Donahue, Neil M.; Downard, Andrew; Dunne, Eimear; Duplissy, Jonathan; Ehrhart, Sebastian; Flagan, Richard C.; Franchin, Alessandro; Guida, Roberto; Hakala, Jani; Hansel, Armin; Heinritzi, Martin; Henschel, Henning; Jokinen, Tuija; Junninen, Heikki; Kajos, Maija; Kangasluoma, Juha; Keskinen, Helmi; Kupc, Agnieszka; Kurten, Theo; Kvashin, Alexander N.; Laaksonen, Ari; Lehtipalo, Katrianne; Leiminger, Markus; Leppae, Johannes; Loukonen, Ville; Makhmutov, Vladimir; Mathot, Serge; McGrath, Matthew J.; Nieminen, Tuomo; Olenius, Tinja; Onnela, Antti; Petaejae, Tuukka; Riccobono, Francesco; Riipinen, Ilona; Rissanen, Matti; Rondo, Linda; Ruuskanen, Taina; Santos, Filipe D.; Sarnela, Nina; Schallhart, Simon; Schnitzhofer, Ralf; Seinfeld, John H.; Simon, Mario; Sipilae, Mikko; Stozhkov, Yuri; Stratmann, Frank; Tome, Antonio; Troestl, Jasmin; Tsagkogeorgas, Georgios; Vaattovaara, Petri; Viisanen, Yrjo; Virtanen, Annele; Vrtala, Aron; Wagner, Paul E.; Weingartner, Ernest; Wex, Heike; Williamson, Christina; Wimmer, Daniela; Ye, Penglin; Yli-Juuti, Taina; Carslaw, Kenneth S.; Kulmala, Markku; Curtius, Joachim; Baltensperger, Urs; Worsnop, Douglas R.; Vehkamaeki, Hanna; Kirkby, JasperNature (London, United Kingdom) (2013), 502 (7471), 359-363CODEN: NATUAS; ISSN:0028-0836. (Nature Publishing Group)Aerosol particle nucleation from trace atm. vapor is thought to provide up to half the global cloud condensation nuclei. Aerosols can cause net climate cooling by scattering sunlight and leading to smaller, but more numerous cloud droplets, making clouds brighter and extending their lifetimes. Atm. aerosols from human activities are thought to have compensated for a large fraction of warming caused by greenhouse gases; however, despite its importance for climate, atm. nucleation is poorly understood. It was recently shown that H2SO4 and NH3 cannot explain particle formation rates obsd. in the lower atm. It is thought amines may enhance nucleation, but until now there has been no direct evidence for amine ternary nucleation under atm. conditions. This work used the CLOUD (cosmics leaving outdoor droplets) chamber at CERN to det. that dimethylamine >3 parts per trillion by vol. can enhance particle formation rates >1000-fold vs. NH3, sufficient to account for atm. obsd. particle formation rates. Mol. anal. of clusters showed that faster nucleation is explained by a base-stabilization mechanism involving acid-amine pairs, which strongly decrease evapn. The ion-induced contribution is generally small, reflecting the high stability of H2SO4-dimethylamine clusters and indicating that galactic cosmic rays exert only a small effect on their formation, except at low overall formation rates. Exptl. measurements were well reproduced by a dynamic model based on quantum chem. calcns. of mol. cluster binding energies without any fitted parameters. Results showed that in regions of the atm. near amine sources, amines and SO2 should be considered when assessing the effect of anthropogenic activity on particle formation.
- 11Kurten, A.; Li, C.; Bianchi, F.; Curtius, J.; Dias, A.; Donahue, N. M.; Duplissy, J.; Flagan, R. C.; Hakala, J.; Jokinen, T. New particle formation in the sulfuric acid-dimethylamine-water system: Reevaluation of CLOUD chamber measurements and comparison to an aerosol nucleation and growth model. Atmospheric Chemistry and Physics 2018, 18, 845– 863, DOI: 10.5194/acp-18-845-2018Google ScholarThere is no corresponding record for this reference.
- 12Rong, H.; Liu, J.; Zhang, Y.; Du, L.; Zhang, X.; Li, Z. Nucleation mechanisms of iodic acid in clean and polluted coastal regions. Chemosphere 2020, 253, 126743, DOI: 10.1016/j.chemosphere.2020.126743Google Scholar12https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXnsFait7s%253D&md5=661c4cafce75612ebd720a3f33d3416bNucleation mechanisms of iodic acid in clean and polluted coastal regionsRong, Hui; Liu, Jiarong; Zhang, Yujia; Du, Lin; Zhang, Xiuhui; Li, ZeshengChemosphere (2020), 253 (), 126743CODEN: CMSHAF; ISSN:0045-6535. (Elsevier Ltd.)In coastal regions, intense bursts of particles are frequently obsd. with high concns. of iodine species, esp. iodic acid (IA). However, the nucleation mechanisms of IA, esp. in polluted environments with high concns. of sulfuric acid (SA) and ammonia (A), remain to be fully established. By quantum chem. calcns. and atm. cluster dynamics code (ACDC) simulations, the self-nucleation of IA in clean coastal regions and that influenced by SA and A in polluted coastal regions are investigated. The results indicate that IA can form stable clusters stabilized by halogen bonds and hydrogen bonds through sequential addn. of IA, and the self-nucleation of IA can instantly produce large amts. of stable clusters when the concn. of IA is high during low tide, which is consistent with the observation that intense particle bursts were linked to high concns. of IA in clean coastal regions. Besides, SA and A can stabilize IA clusters by the formation of more halogen bonds and hydrogen bonds as well as proton transfers, and the binary nucleation of IA-SA/A rather than the self-nucleation of IA appears to be the dominant pathways in polluted coastal regions, esp. in winter. These new insights are helpful to understand the mechanisms of new particle formation induced by IA in clean and polluted coastal regions.
- 13He, X.-C.; Tham, Y. J.; Dada, L.; Wang, M.; Finkenzeller, H.; Stolzenburg, D.; Iyer, S.; Simon, M.; Kürten, A.; Shen, J. Role of iodine oxoacids in atmospheric aerosol nucleation. Science 2021, 371, 589– 595, DOI: 10.1126/science.abe0298Google Scholar13https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXjsFCrtr8%253D&md5=f7b67dfba68a1d92579197f81529327bRole of iodine oxoacids in atmospheric aerosol nucleationHe, Xu-Cheng; Tham, Yee Jun; Dada, Lubna; Wang, Mingyi; Finkenzeller, Henning; Stolzenburg, Dominik; Iyer, Siddharth; Simon, Mario; Kurten, Andreas; Shen, Jiali; Rorup, Birte; Rissanen, Matti; Schobesberger, Siegfried; Baalbaki, Rima; Wang, Dongyu S.; Koenig, Theodore K.; Jokinen, Tuija; Sarnela, Nina; Beck, Lisa J.; Almeida, Joao; Amanatidis, Stavros; Amorim, Antonio; Ataei, Farnoush; Baccarini, Andrea; Bertozzi, Barbara; Bianchi, Federico; Brilke, Sophia; Caudillo, Lucia; Chen, Dexian; Chiu, Randall; Chu, Biwu; Dias, Antonio; Ding, Aijun; Dommen, Josef; Duplissy, Jonathan; El Haddad, Imad; Gonzalez Carracedo, Loic; Granzin, Manuel; Hansel, Armin; Heinritzi, Martin; Hofbauer, Victoria; Junninen, Heikki; Kangasluoma, Juha; Kemppainen, Deniz; Kim, Changhyuk; Kong, Weimeng; Krechmer, Jordan E.; Kvashin, Aleksander; Laitinen, Totti; Lamkaddam, Houssni; Lee, Chuan Ping; Lehtipalo, Katrianne; Leiminger, Markus; Li, Zijun; Makhmutov, Vladimir; Manninen, Hanna E.; Marie, Guillaume; Marten, Ruby; Mathot, Serge; Mauldin, Roy L.; Mentler, Bernhard; Mohler, Ottmar; Muller, Tatjana; Nie, Wei; Onnela, Antti; Petaja, Tuukka; Pfeifer, Joschka; Philippov, Maxim; Ranjithkumar, Ananth; Saiz-Lopez, Alfonso; Salma, Imre; Scholz, Wiebke; Schuchmann, Simone; Schulze, Benjamin; Steiner, Gerhard; Stozhkov, Yuri; Tauber, Christian; Tome, Antonio; Thakur, Roseline C.; Vaisanen, Olli; Vazquez-Pufleau, Miguel; Wagner, Andrea C.; Wang, Yonghong; Weber, Stefan K.; Winkler, Paul M.; Wu, Yusheng; Xiao, Mao; Yan, Chao; Ye, Qing; Ylisirnio, Arttu; Zauner-Wieczorek, Marcel; Zha, Qiaozhi; Zhou, Putian; Flagan, Richard C.; Curtius, Joachim; Baltensperger, Urs; Kulmala, Markku; Kerminen, Veli-Matti; Kurten, Theo; Donahue, Neil M.; Volkamer, Rainer; Kirkby, Jasper; Worsnop, Douglas R.; Sipila, MikkoScience (Washington, DC, United States) (2021), 371 (6529), 589-595CODEN: SCIEAS; ISSN:1095-9203. (American Association for the Advancement of Science)Iodic acid (HIO3) is known to form aerosol particles in coastal marine regions, but predicted nucleation and growth rates are lacking. Using the CERN CLOUD (Cosmics Leaving Outdoor Droplets) chamber, we find that the nucleation rates of HIO3 particles are rapid, even exceeding sulfuric acid-ammonia rates under similar conditions. We also find that ion-induced nucleation involves IO3- and the sequential addn. of HIO3 and that it proceeds at the kinetic limit below +10°C. In contrast, neutral nucleation involves the repeated sequential addn. of iodous acid (HIO2) followed by HIO3, showing that HIO2 plays a key stabilizing role. Freshly formed particles are composed almost entirely of HIO3, which drives rapid particle growth at the kinetic limit. Our measurements indicate that iodine oxoacid particle formation can compete with sulfuric acid in pristine regions of the atm.
- 14Baccarini, A.; Karlsson, L.; Dommen, J.; Duplessis, P.; Vullers, J.; Brooks, I. M.; Saiz-Lopez, A.; Salter, M.; Tjernstrom, M.; Baltensperger, U. Frequent new particle formation over the high Arctic pack ice by enhanced iodine emissions. Nat. Commun. 2020, 11, 4924, DOI: 10.1038/s41467-020-19533-yGoogle Scholar14https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXhvFyqt7jM&md5=8ce5a303b90658edab0b26febba072cbFrequent new particle formation over the high Arctic pack ice by enhanced iodine emissionsBaccarini, Andrea; Karlsson, Linn; Dommen, Josef; Duplessis, Patrick; Vullers, Jutta; Brooks, Ian M.; Saiz-Lopez, Alfonso; Salter, Matthew; Tjernstrom, Michael; Baltensperger, Urs; Zieger, Paul; Schmale, JuliaNature Communications (2020), 11 (1), 4924CODEN: NCAOBW; ISSN:2041-1723. (Nature Research)In the central Arctic Ocean the formation of clouds and their properties are sensitive to the availability of cloud condensation nuclei (CCN). The vapors responsible for new particle formation (NPF), potentially leading to CCN, have remained unidentified since the first aerosol measurements in 1991. Here, we report that all the obsd. NPF events from the Arctic Ocean 2018 expedition are driven by iodic acid with little contribution from sulfuric acid. Iodic acid largely explains the growth of ultrafine particles (UFP) in most events. The iodic acid concn. increases significantly from summer towards autumn, possibly linked to the ocean freeze-up and a seasonal rise in ozone. This leads to a one order of magnitude higher UFP concn. in autumn. Measurements of cloud residuals suggest that particles smaller than 30 nm in diam. can activate as CCN. Therefore, iodine NPF has the potential to influence cloud properties over the Arctic Ocean.
- 15Jimenez, J. L.; Canagaratna, M. R.; Donahue, N. M.; Prevot, A. S. H.; Zhang, Q.; Kroll, J. H.; DeCarlo, P. F.; Allan, J. D.; Coe, H.; Ng, N. L. Evolution of Organic Aerosols in the Atmosphere. Science (New York, N.Y.) 2009, 326, 1525– 9, DOI: 10.1126/science.1180353Google Scholar15https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhsFensbjE&md5=dd5505995c2591c0540180493b0020eeEvolution of Organic Aerosols in the AtmosphereJimenez, J. L.; Canagaratna, M. R.; Donahue, N. M.; Prevot, A. S. H.; Zhang, Q.; Kroll, J. H.; DeCarlo, P. F.; Allan, J. D.; Coe, H.; Ng, N. L.; Aiken, A. C.; Docherty, K. S.; Ulbrich, I. M.; Grieshop, A. P.; Robinson, A. L.; Duplissy, J.; Smith, J. D.; Wilson, K. R.; Lanz, V. A.; Hueglin, C.; Sun, Y. L.; Tian, J.; Laaksonen, A.; Raatikainen, T.; Rautiainen, J.; Vaattovaara, P.; Ehn, M.; Kulmala, M.; Tomlinson, J. M.; Collins, D. R.; Cubison, M. J.; Dunlea, J.; Huffman, J. A.; Onasch, T. B.; Alfarra, M. R.; Williams, P. I.; Bower, K.; Kondo, Y.; Schneider, J.; Drewnick, F.; Borrmann, S.; Weimer, S.; Demerjian, K.; Salcedo, D.; Cottrell, L.; Griffin, R.; Takami, A.; Miyoshi, T.; Hatakeyama, S.; Shimono, A.; Sun, J. Y.; Zhang, Y. M.; Dzepina, K.; Kimmel, J. R.; Sueper, D.; Jayne, J. T.; Herndon, S. C.; Trimborn, A. M.; Williams, L. R.; Wood, E. C.; Middlebrook, A. M.; Kolb, C. E.; Baltensperger, U.; Worsnop, D. R.Science (Washington, DC, United States) (2009), 326 (5959), 1525-1529CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)Org. aerosol (OA) particles affect climate forcing and human health, but their sources and evolution are poorly characterized. A unifying model framework describing the atm. evolution of OA which is constrained by high time resolved measurements of its compn., volatility, and oxidn. state is presented. OA and OA precursor gases evolve by becoming increasingly oxidized, less volatile, and more hygroscopic, leading to the formation of oxygenated org. aerosols (OOA), with concns. comparable to those of SO42- aerosols throughout the Northern Hemisphere. This model framework captures the dynamic aging behavior obsd. in the atm. and lab.; it serves as a basis to improve regional and global model parameterizations.
- 16Petaja, T.; Tabakova, K.; Manninen, A.; Ezhova, E.; O’Connor, E.; Moisseev, D.; Sinclair, V. A.; Backman, J.; Levula, J.; Luoma, K. Influence of biogenic emissions from boreal forests on aerosol-cloud interactions. Nature Geoscience 2022, 15, 42– 47, DOI: 10.1038/s41561-021-00876-0Google ScholarThere is no corresponding record for this reference.
- 17Tröstl, J.; Chuang, W.; Gordon, H.; Heinritzi, M.; Yan, C.; Molteni, U.; Ahlm, L.; Frege, C.; Bianchi, F.; Wagner, R. The role of low-volatility organic compounds in initial particle growth in the atmosphere. Nature 2016, 533, 527– 531, DOI: 10.1038/nature18271Google Scholar17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28Xoslyrur4%253D&md5=7e09ef77e9daeb1f54cb52383f4feb68The role of low-volatility organic compounds in initial particle growth in the atmosphereTrostl, Jasmin; Chuang, Wayne K.; Gordon, Hamish; Heinritzi, Martin; Yan, Chao; Molteni, Ugo; Ahlm, Lars; Frege, Carla; Bianchi, Federico; Wagner, Robert; Simon, Mario; Lehtipalo, Katrianne; Williamson, Christina; Craven, Jill S.; Duplissy, Jonathan; Adamov, Alexey; Almeida, Joao; Bernhammer, Anne-Kathrin; Breitenlechner, Martin; Brilke, Sophia; Dias, Antonio; Ehrhart, Sebastian; Flagan, Richard C.; Franchin, Alessandro; Fuchs, Claudia; Guida, Roberto; Gysel, Martin; Hansel, Armin; Hoyle, Christopher R.; Jokinen, Tuija; Junninen, Heikki; Kangasluoma, Juha; Keskinen, Helmi; Kim, Jaeseok; Krapf, Manuel; Kurten, Andreas; Laaksonen, Ari; Lawler, Michael; Leiminger, Markus; Mathot, Serge; Mohler, Ottmar; Nieminen, Tuomo; Onnela, Antti; Petaja, Tuukka; Piel, Felix M.; Miettinen, Pasi; Rissanen, Matti P.; Rondo, Linda; Sarnela, Nina; Schobesberger, Siegfried; Sengupta, Kamalika; Sipila, Mikko; Smith, James N.; Steiner, Gerhard; Tome, Antonio; Virtanen, Annele; Wagner, Andrea C.; Weingartner, Ernest; Wimmer, Daniela; Winkler, Paul M.; Ye, Penglin; Carslaw, Kenneth S.; Curtius, Joachim; Dommen, Josef; Kirkby, Jasper; Kulmala, Markku; Riipinen, Ilona; Worsnop, Douglas R.; Donahue, Neil M.; Baltensperger, UrsNature (London, United Kingdom) (2016), 533 (7604), 527-531CODEN: NATUAS; ISSN:0028-0836. (Nature Publishing Group)About half of present-day cloud condensation nuclei originate from atm. nucleation, frequently appearing as a burst of new particles near midday. Atm. observations show that the growth rate of new particles often accelerates when the diam. of the particles is between one and ten nanometers. In this crit. size range, new particles are most likely to be lost by coagulation with pre-existing particles, thereby failing to form new cloud condensation nuclei that are typically 50 to 100 nm across. Sulfuric acid vapor is often involved in nucleation but is too scarce to explain most subsequent growth, leaving org. vapors as the most plausible alternative, at least in the planetary boundary layer. Although recent studies predict that low-volatility org. vapors contribute during initial growth, direct evidence has been lacking. The accelerating growth may result from increased photolytic prodn. of condensable org. species in the afternoon, and the presence of a possible Kelvin (curvature) effect, which inhibits org. vapor condensation on the smallest particles (the nano-Kohler theory), has so far remained ambiguous. The authors presented expts. performed in a large chamber under atm. conditions that investigate the role of org. vapors in the initial growth of nucleated org. particles in the absence of inorg. acids and bases such as sulfuric acid or ammonia and amines, resp. Using data from the same set of expts., it has been shown that org. vapors alone can drive nucleation. They focused on the growth of nucleated particles and find that the org. vapors that drive initial growth have extremely low volatilities (satn. concn. less than 10-4.5 micrograms per cubic meter). As the particles increase in size and the Kelvin barrier falls, subsequent growth is primarily due to more abundant org. vapors of slightly higher volatility (satn. concns. of 10-4.5 to 10-0.5 micrograms per cubic meter). They present a particle growth model that quant. reproduces measurements. They implement a parameterization of the first steps of growth in a global aerosol model and find that concns. of atm. cloud concn. nuclei can change substantially in response, i.e., by up to 50 per cent in comparison with previously assumed growth rate parameterizations.
- 18Crounse, J.; Nielsen, L.; Jørgensen, S.; Kjaergaard, H.; Wennberg, P. Autoxidation of Organic Compounds in the Atmosphere. JOURNAL OF PHYSICAL. CHEMISTRY LETTERS 2013, 4, 3513– 3520, DOI: 10.1021/jz4019207Google ScholarThere is no corresponding record for this reference.
- 19Bianchi, F.; Kurtén, T.; Riva, M.; Mohr, C.; Rissanen, M.; Roldin, P.; Berndt, T.; Crounse, J.; Wennberg, P.; Mentel, T.; Wildt, J.; Junninen, H.; Jokinen, T.; Kulmala, M.; Worsnop, D.; Thornton, J.; Donahue, N.; Kjaergaard, H.; Ehn, M. Highly Oxygenated Molecules (HOM) from Gas-Phase Autoxidation Involving Organic Peroxy Radicals: A Key Contributor to Atmospheric Aerosol. Chem. Rev. 2019, 119, 3472– 3509, DOI: 10.1021/acs.chemrev.8b00395Google Scholar19https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXjsFSrsb0%253D&md5=860f5dce6415fbbdee9f45b7afeff559Highly Oxygenated Molecules (HOM) from Gas-Phase Autoxidation Involving Organic Peroxy Radicals: A Key Contributor to Atmospheric AerosolBianchi, Federico; Kurten, Theo; Riva, Matthieu; Mohr, Claudia; Rissanen, Matti P.; Roldin, Pontus; Berndt, Torsten; Crounse, John D.; Wennberg, Paul O.; Mentel, Thomas F.; Wildt, Jurgen; Junninen, Heikki; Jokinen, Tuija; Kulmala, Markku; Worsnop, Douglas R.; Thornton, Joel A.; Donahue, Neil; Kjaergaard, Henrik G.; Ehn, MikaelChemical Reviews (Washington, DC, United States) (2019), 119 (6), 3472-3509CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review which defines highly oxygenated org. mols. (HOM) formed in the atm. via auto-oxidn. involving peroxy radicals arising from volatile org. compds. describing currently available techniques for their identification/quantification, followed by a summary of the current knowledge on their formation mechanisms and physicochem. properties, is given. Major aims are to provide a common frame for the currently quite fragmented literature on HOM studies and highlighting existing gaps, and suggesting directions for future HOM research. Topics discussed include: introduction; HOM background (defining key concepts, HOM in relation to other classification schemes, historical naming conventions); HOM detection (gas and particle phases, uncertainties and anal. challenges of HOM detection); HOM formation mechanisms (auto-oxidn. involving peroxy radical as HOM source, bimol. RO2 reactions, factors affecting HOM formation); HOM properties and atm. fate (physicochem. properties, removal mechanisms); HOM atm. observations and impact (ambient HOM observation, atm. impact); and summary and perspectives.
- 20Roldin, P.; Ehn, M.; Kurten, T.; Olenius, T.; Rissanen, M. P.; Sarnela, N.; Elm, J.; Rantala, P.; Hao, L.; Hyttinen, N. The role of highly oxygenated organic molecules in the Boreal aerosol-cloud-climate system. Nat. Commun. 2019, 10, 1– 15, DOI: 10.1038/s41467-019-12338-8Google ScholarThere is no corresponding record for this reference.
- 21Xavier, C.; de jonge, R. W.; Jokinen, T.; Beck, L.; Sipilä, M.; Olenius, T.; Roldin, P. Role of Iodine-Assisted Aerosol Particle Formation in Antarctica. Environ. Sci. Technol. 2024, 58, 7314– 7324, DOI: 10.1021/acs.est.3c09103Google ScholarThere is no corresponding record for this reference.
- 22Brean, J.; Dall’Osto, M.; Simo, R.; Shi, Z.; Beddows, D. C. S.; Harrison, R. M. Open ocean and coastal new particle formation from sulfuric acid and amines around the Antarctic Peninsula. Nature Geoscience 2021, 14, 383– 388, DOI: 10.1038/s41561-021-00751-yGoogle ScholarThere is no corresponding record for this reference.
- 23Jokinen, T.; Sipila, M.; Kontkanen, J.; Vakkari, V.; Tisler, P.; Duplissy, E.-M.; Junninen, H.; Kangasluoma, J.; Manninen, H. E.; Petaja, T. Ion-induced sulfuric acid-ammonia nucleation drives particle formation in coastal Antarctica. Science Advances 2018, 4, eaat9744, DOI: 10.1126/sciadv.aat9744Google ScholarThere is no corresponding record for this reference.
- 24Lee, H.; Lee, K.; Lunder, C.; Krejci, R.; Aas, W.; Jiyeon, P.; Park, K.-t.; Lee, B.; Yoon, Y.-J.; Park, K. Atmospheric new particle formation characteristics in the Arctic as measured at Mount Zeppelin, Svalbard, from 2016 to 2018. Atmospheric Chem. Phys. 2020, 20 (21), 13425– 13441Google ScholarThere is no corresponding record for this reference.
- 25Kecorius, S.; Vogl, T.; Paasonen, P.; Lampilahti, J.; Rothenberg, D.; Wex, H.; Zeppenfeld, S.; van Pinxteren, M.; Hartmann, M.; Henning, S. New particle formation and its effect on cloud condensation nuclei abundance in the summer Arctic: a case study in the Fram Strait and Barents Sea. Atmospheric Chemistry and Physics 2019, 19, 14339– 14364, DOI: 10.5194/acp-19-14339-2019Google ScholarThere is no corresponding record for this reference.
- 26DallÓsto, M.; Geels, C.; Beddows, D. C. S.; Boertmann, D.; Lange, R.; Nøjgaard, J. K.; Harrison, R. M.; Simo, R.; Skov, H.; Massling, A. Regions of open water and melting sea ice drive new particle formation in North East Greenland. Sci. Rep. 2018, 8, 6109, DOI: 10.1038/s41598-018-24426-8Google ScholarThere is no corresponding record for this reference.
- 27Mäkelä, J.; Yli-Koivisto, S.; Hiltunen, V.; Seidl, W.; Swietlicki, E.; Teinilä, K.; Sillanpää, M.; Koponen, I.; Paatero, J.; Rosman, K.; Hämeri, K. Chemical composition of aerosol during particle formation events in boreal forest. Tellus B 2001, 53, 380– 393, DOI: 10.1034/j.1600-0889.2001.530405.xGoogle Scholar27https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXns1Gktrs%253D&md5=ce371260e6101c15c4bbaac0a5473bb1Chemical composition of aerosol during particle formation events in boreal forestMakela, J. M.; Yli-Koivisto, S.; Hiltunen, V.; Seidl, W.; Swietlicki, E.; Teinila, K.; Sillanpaa, M.; Koponen, I. K.; Paatero, J.; Rosman, K.; Hameri, K.Tellus, Series B: Chemical and Physical Meteorology (2001), 53B (4), 380-393CODEN: TSBMD7; ISSN:0280-6509. (Munksgaard International Publishers Ltd.)Size-segregated chem. aerosol anal. of 5 integrated samples was performed for atm. aerosols during new particle formation events in spring 1999 during the BIOFOR 3 measurement campaign at a boreal forest site in southern Finland. Aerosol samples collected by a cascade low-pressure impactor were taken selectively to distinguish particle formation event aerosols from non-event aerosols. The division into event and non-event cases was done in-situ in the field, based on the online submicron no. size distribution. Results of chem. ionic compn. of particles showed only small differences between event and non-event sample sets. Event samples had lower concns. of total SO42- and NH4+ and light dicarboxylic acids, e.g., oxalate, malonate, and succinate. In event samples, nucleation mode particle MSA (methane sulfonic acid) was present in concns. exceeding those obsd. in non-event samples; however, at larger particle sizes, sample sets contained rather similar MSA concns. The most significant difference between event and non-event sets was obsd. for dimethylammonium, the ionic component of dimethylamine ((CH3)2NH), which seemed to be present in the particle phase during particle formation periods and/or during subsequent particle growth. The abs. event sample dimethylamine concns. were >30-fold greater than non-event concns. in the accumulation mode size range. The non-event back-up filter stage for sub-30 nm particles contained more dimethylamine than event samples. This fractionation is probably a condensation artifact from impactor sampling. A simple mass balance est. was performed to evaluate the quality and consistency of results for overall mass concn.
- 28Hemmilä, M.; Hellén, H.; Virkkula, A.; Makkonen, U.; Praplan, A.; Kontkanen, J.; Ahonen, L.; Kulmala, M.; Hakola, H. Amines in boreal forest air at SMEAR II station in Finland. Atmospheric Chemistry and Physics 2018, 18, 6367– 6380, DOI: 10.5194/acp-18-6367-2018Google ScholarThere is no corresponding record for this reference.
- 29Lawler, M.; Rissanen, M.; Ehn, M.; Mauldin, R.; Sarnela, N.; Sipilä, M.; Smith, J. Evidence for Diverse Biogeochemical Drivers of Boreal Forest New Particle Formation. Geophys. Res. Lett. 2018, 45, 2038– 2046, DOI: 10.1002/2017GL076394Google ScholarThere is no corresponding record for this reference.
- 30Sogacheva, L.; Saukkonen, L.; Nilsson, E.; Dal Maso, M.; Schultz, D.; de Leeuw, G.; Kulmala, M. New aerosol particle formation in different synoptic situations at Hyytiälä, Southern Finland. Tellus B 2022, 60, 485– 494, DOI: 10.1111/j.1600-0889.2008.00364.xGoogle ScholarThere is no corresponding record for this reference.
- 31Nieminen, T.; Yli-Juuti, T.; Manninen, H. E.; Petäjä, T.; Kerminen, V.-M.; Kulmala, M. Technical note: New particle formation event forecasts during PEGASOS-Zeppelin Northern mission 2013 in Hyytiälä, Finland. Atmospheric Chemistry and Physics 2015, 15, 12385– 12396, DOI: 10.5194/acp-15-12385-2015Google ScholarThere is no corresponding record for this reference.
- 32Öström, E.; Putian, Z.; Schurgers, G.; Mishurov, M.; Kivekäs, N.; Lihavainen, H.; Ehn, M.; Rissanen, M. P.; Kurtén, T.; Boy, M.; Swietlicki, E.; Roldin, P. Modeling the role of highly oxidized multifunctional organic molecules for the growth of new particles over the boreal forest region. Atmospheric Chemistry and Physics 2017, 17, 8887– 8901, DOI: 10.5194/acp-17-8887-2017Google ScholarThere is no corresponding record for this reference.
- 33Jenkin, M. E.; Saunders, S. M.; Pilling, M. J. The tropospheric degradation of volatile organic compounds: a protocol for mechanism development. Atmos. Environ. 1997, 31, 81– 104, DOI: 10.1016/S1352-2310(96)00105-7Google Scholar33https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XntVOnt78%253D&md5=44dc875e8e1a9456ed2b570393d450c6The tropospheric degradation of volatile organic compounds: a protocol for mechanism developmentJenkin, Michael E.; Saunders, Sandra M.; Pilling, Michael J.Atmospheric Environment (1996), 31 (1), 81-104CODEN: AENVEQ; ISSN:1352-2310. (Elsevier)Kinetic and mechanistic data relevant to the tropospheric oxidn. of volatile org. compds. (VOCs) were used to define a series of rules for the construction of detailed degrdn. schemes for use in numerical models. These rules are intended to apply to the treatment of a wide range of non-arom. hydrocarbons and oxygenated and chlorinated VOCs, and are currently used to provide an up-to-date mechanism describing the degrdn. of a range of VOCs, and the formation of secondary oxidants, for use in a model of the boundary layer over Europe. The schemes constructed using this protocol are applicable, however, to a wide range of ambient conditions, and may be employed in models of urban, rural, or remote tropospheric environments, or for the simulation of secondary pollutant formation for a range of NOx or VOC emission scenarios. These schemes are believed to be particularly appropriate for comparative assessments of the formation of oxidants, such as ozone, from the degrdn. of org. compds. The protocol is divided into a series of subsections dealing with initiation reactions, the reactions of the radical intermediates and the further degrdn. of first and subsequent generation products. The present work draws heavily on previous reviews and evaluations of data relevant to tropospheric chem. Where necessary, however, existing recommendations are adapted, or new rules are defined, to reflect recent improvements in the database, particularly with regard to the treatment of peroxy radical (RO2) reactions for which there have been major advances, even since comparatively recent reviews. The present protocol aims to take into consideration work available in the open literature up to the end of 1994, and some further studies known by the authors, which were under review at that time. A major disadvantage of explicit chem. mechanisms is the very large no. of reactions potentially generated, if a series of rules is rigorously applied. The protocol aims to limit the no. of reactions in a degrdn. scheme by applying a degree of strategic simplication, while maintaining the essential features of the chem. These simplication measures are described, and their influence is demonstrated and discussed. The resultant mechanisms are believed to provide a suitable starting point for the generation of reduced chem. mechanisms.
- 34Jenkin, M. E.; Saunders, S. M.; Wagner, V.; Pilling, M. J. Protocol for the development of the Master Chemical Mechanism, MCM v3 (Part B): tropospheric degradation of aromatic volatile organic compounds. Atmospheric Chemistry and Physics 2003, 3, 181– 193, DOI: 10.5194/acp-3-181-2003Google Scholar34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXns1Sntrw%253D&md5=cbb782c7d9ee66115b9c13f14758d095Protocol for the development of the master chemical mechanism, MCM v3 (part B): tropospheric degradation of aromatic volatile organic compoundsJenkin, M. E.; Saunders, S. M.; Wagner, V.; Pilling, M. J.Atmospheric Chemistry and Physics (2003), 3 (1), 181-193CODEN: ACPTCE; ISSN:1680-7324. (European Geophysical Society)Kinetic and mechanistic data relevant to the tropospheric degrdn. of arom. volatile org. compds. (VOC) were used to define a mechanism development protocol, which was used to construct degrdn. schemes for 18 arom. VOC as part of version 3 of the Master Chem. Mechanism (MCM v3). This is complementary to the treatment of 107 nonarom. VOC, presented in a companion paper. The protocol is divided into subsections describing initiation reactions, the degrdn. chem. to 1st generation products via a no. of competitive routes, and the further degrdn. of 1st and subsequent generation products. Emphasis is placed on describing where the treatment differs from that applied to the nonarom. VOC. The protocol is based on work available in the open literature up to the beginning of 2001, and some other studies known by the authors which were under review at the time. Photochem. Ozone Creation Potentials (POCP) were calcd. for the 18 arom. VOC in MCM v3 for idealized conditions appropriate to north-west Europe, using a photochem. trajectory model. The POCP values provide a measure of the relative ozone forming abilities of the VOC. These show distinct differences from POCP values calcd. previously for the aroms., using earlier versions of the MCM, and reasons for these differences are discussed.
- 35Braeuer, P.; Tilgner, A.; Wolke, R.; Herrmann, H. Mechanism development and modelling of tropospheric multiphase halogen chemistry: The CAPRAM Halogen Module 2.0 (HM2). JOURNAL OF ATMOSPHERIC CHEMISTRY 2013, 70, 19– 52, DOI: 10.1007/s10874-013-9249-6Google ScholarThere is no corresponding record for this reference.
- 36Wu, R.; Wang, S.; Wang, L. A New Mechanism for The Atmospheric Oxidation of Dimethyl Sulfide. The Importance of Intramolecular Hydrogen Shift in CH3SCH2OO Radical. journal of physical chemistry. A 2015, 119, 112, DOI: 10.1021/jp511616jGoogle ScholarThere is no corresponding record for this reference.
- 37Berndt, T.; Scholz, W.; Mentler, B.; Fischer, L.; Hoffmann, E.; Tilgner, A.; Hyttinen, N.; Prisle, N. L.; Hansel, A.; Herrmann, H. Fast Peroxy Radical Isomerization and OH Recycling in the Reaction of OH Radicals with Dimethyl Sulfide. J. Phys. Chem. Lett. 2019, 10, 6478, DOI: 10.1021/acs.jpclett.9b02567Google Scholar37https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXhvFWltrjP&md5=0d91ffa9cadc1365ffbfb0a48fe0373bFast Peroxy Radical Isomerization and OH Recycling in the Reaction of OH Radicals with Dimethyl SulfideBerndt, T.; Scholz, W.; Mentler, B.; Fischer, L.; Hoffmann, E. H.; Tilgner, A.; Hyttinen, N.; Prisle, N. L.; Hansel, A.; Herrmann, H.Journal of Physical Chemistry Letters (2019), 10 (21), 6478-6483CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)Di-Me sulfide (DMS), produced by marine organisms, represents the most abundant, biogenic sulfur emission into the Earth's atm. The gas-phase degrdn. of DMS is mainly initiated by the reaction with the OH radical forming first CH3SCH2O2 radicals from the dominant H-abstraction channel. It is exptl. shown that these peroxy radicals undergo a two-step isomerization process finally forming a product consistent with the formula HOOCH2SCHO. The isomerization process is accompanied by OH recycling. The rate-limiting first isomerization step, CH3SCH2O2 → CH2SCH2OOH, followed by O2 addn., proceeds with k = (0.23 ± 0.12) s-1 at 295 ± 2 K. Competing bimol. CH3SCH2O2 reactions with NO, HO2, or RO2 radicals are less important for trace-gas conditions over the oceans. Results of atm. chem. simulations demonstrate the predominance (≥95%) of CH3SCH2O2 isomerization. The rapid peroxy radical isomerization, not yet considered in models, substantially changes the understanding of DMS's degrdn. processes in the atm.
- 38Veres, P. R.; Neuman, J. A.; Bertram, T. H.; Assaf, E.; Wolfe, G. M.; Williamson, C. J.; Weinzierl, B.; Tilmes, S.; Thompson, C. R.; Thames, A. B.; Schroder, J. C.; Saiz-Lopez, A.; Rollins, A. W.; Roberts, J. M.; Price, D.; Peischl, J.; Nault, B. A.; Møller, K. H.; Miller, D. O.; Meinardi, S.; Li, Q.; Lamarque, J.-F.; Kupc, A.; Kjaergaard, H. G.; Kinnison, D.; Jimenez, J. L.; Jernigan, C. M.; Hornbrook, R. S.; Hills, A.; Dollner, M.; Day, D. A.; Cuevas, C. A.; Campuzano-Jost, P.; Burkholder, J.; Bui, T. P.; Brune, W. H.; Brown, S. S.; Brock, C. A.; Bourgeois, I.; Blake, D. R.; Apel, E. C.; Ryerson, T. B. Global airborne sampling reveals a previously unobserved dimethyl sulfide oxidation mechanism in the marine atmosphere. Proc. Natl. Acad. Sci. U. S. A. 2020, 117, 4505– 4510, DOI: 10.1073/pnas.1919344117Google Scholar38https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXksVSqsr0%253D&md5=0d957baa667a2543cd9a4f3c467790a5Global airborne sampling reveals a previously unobserved dimethyl sulfide oxidation mechanism in the marine atmosphereVeres, Patrick R.; Neuman, J. Andrew; Bertram, Timothy H.; Assaf, Emmanuel; Wolfe, Glenn M.; Williamson, Christina J.; Weinzierl, Bernadett; Tilmes, Simone; Thompson, Chelsea R.; Thames, Alexander B.; Schroder, Jason C.; Saiz-Lopez, Alfonso; Rollins, Andrew W.; Roberts, James M.; Price, Derek; Peischl, Jeff; Nault, Benjamin A.; Moeller, Kristian H.; Miller, David O.; Meinardi, Simone; Li, Qinyi; Lamarque, Jean-Francois; Kupc, Agnieszka; Kjaergaard, Henrik G.; Kinnison, Douglas; Jimenez, Jose L.; Jernigan, Christopher M.; Hornbrook, Rebecca S.; Hills, Alan; Dollner, Maximilian; Day, Douglas A.; Cuevas, Carlos A.; Campuzano-Jost, Pedro; Burkholder, James; Bui, T. Paul; Brune, William H.; Brown, Steven S.; Brock, Charles A.; Bourgeois, Ilann; Blake, Donald R.; Apel, Eric C.; Ryerson, Thomas B.Proceedings of the National Academy of Sciences of the United States of America (2020), 117 (9), 4505-4510CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)Di-Me sulfide (DMS), emitted from the oceans, is the most abundant biol. source of sulfur to the marine atm. Atm. DMS is oxidized to condensable products that form secondary aerosols that affect Earth's radiative balance by scattering solar radiation and serving as cloud condensation nuclei. We report the atm. discovery of a previously unquantified DMS oxidn. product, hydroperoxymethyl thioformate (HPMTF, HOOCH2SCHO), identified through global-scale airborne observations that demonstrate it to be a major reservoir of marine sulfur. Observationally constrained model results show that more than 30% of oceanic DMS emitted to the atm. forms HPMTF. Coincident particle measurements suggest a strong link between HPMTF concn. and new particle formation and growth. Analyses of these observations show that HPMTF chem. must be included in atm. models to improve representation of key linkages between the biogeochem. of the ocean, marine aerosol formation and growth, and their combined effects on climate.
- 39Hari, P.; Kulmala, M. Station for Measuring Ecosystem-Atmosphere Relations (SMEAR II). Boreal Environment Research 2005, 10, 315– 322Google Scholar39https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXht1GhurvP&md5=fbd8465f76a998b3a0958c12089db374Station for Measuring Ecosystem-Atmosphere Relations (SMEAR II)Hari, Pertti; Kulmala, MarkkuBoreal Environment Research (2005), 10 (5), 315-322CODEN: BEREF7; ISSN:1239-6095. (Finnish Environment Institute)Here we present the ongoing SMEAR (Station for Measuring Forest Ecosystem-Atm. Relations) research program and also related future views. The main idea of SMEAR-type infrastructures is continuous, comprehensive measurements of fluxes, storages and concns. in the land ecosystem-atm. continuum. The major coupling mechanisms between atm. and land surface are the fluxes of energy, momentum, water, carbon dioxide, atm. trace gases and atm. aerosols. Understanding of couplings and feedbacks is the basis for the prediction of changes in the system formed by atm., vegetation and soil. A better quantification of the agents that cause climate change, as well as the emissions and removals of species, will provide more accurate projections of future atm. compn. and hence climate.
- 40Neefjes, I.; Laapas, M.; Liu, Y.; Médus, E.; Miettunen, E.; Ahonen, L.; Quéléver, L.; Aalto, J.; Bäck, J.; Kerminen, V.-M.; Lamplahti, J.; Luoma, K.; Maki, M.; Mammarella, I.; Petäjä, T.; Räty, M.; Sarnela, N.; Ylivinkka, I.; Hakala, S.; Lintunen, A. 25 years of atmospheric and ecosystem measurements in a boreal forestSeasonal variation and responses to warm and dry years. Boreal Environment Research 2022, 27, 1– 31Google ScholarThere is no corresponding record for this reference.
- 41Stein, A. F.; Draxler, R. R.; Rolph, G. D.; Stunder, B. J. B.; Cohen, M. D.; Ngan, F. NOAAs HYSPLIT Atmospheric Transport and Dispersion Modeling System. Bulletin of the American Meteorological Society 2015, 96, 2059– 2077, DOI: 10.1175/BAMS-D-14-00110.1Google ScholarThere is no corresponding record for this reference.
- 42Rolph, G.; Stein, A.; Stunder, B. Real-time Environmental Applications and Display sYstem: READY. Environmental Modelling & Software 2017, 95, 210– 228, DOI: 10.1016/j.envsoft.2017.06.025Google ScholarThere is no corresponding record for this reference.
- 43Lennartz, S. T.; Marandino, C. A.; von Hobe, M.; Cortes, P.; Quack, B.; Simo, R.; Booge, D.; Pozzer, A.; Steinhoff, T.; Arevalo-Martinez, D. L. Direct oceanic emissions unlikely to account for the missing source of atmospheric carbonyl sulfide. Atmospheric Chemistry and Physics 2017, 17, 385– 402, DOI: 10.5194/acp-17-385-2017Google ScholarThere is no corresponding record for this reference.
- 44Ziska, F.; Quack, B.; Abrahamsson, K.; Archer, S. D.; Atlas, E.; Bell, T.; Butler, J. H.; Carpenter, L. J.; Jones, C. E.; Harris, N. R. P. Global sea-to-air flux climatology for bromoform, dibromomethane and methyl iodide. Atmospheric Chemistry and Physics Discussions 2013, 13, 8915– 8934, DOI: 10.5194/acp-13-8915-2013Google Scholar44https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhslehtrbL&md5=d13f982b82d94c732ab4dc017908098cGlobal sea-to-air flux climatology for bromoform, dibromomethane and methyl iodideZiska, F.; Quack, B.; Abrahamsson, K.; Archer, S. D.; Atlas, E.; Bell, T.; Butler, J. H.; Carpenter, L. J.; Jones, C. E.; Harris, N. R. P.; Hepach, H.; Heumann, K. G.; Hughes, C.; Kuss, J.; Krueger, K.; Liss, P.; Moore, R. M.; Orlikowska, A.; Raimund, S.; Reeves, C. E.; Reifenhaeuser, W.; Robinson, A. D.; Schall, C.; Tanhua, T.; Tegtmeier, S.; Turner, S.; Wang, L.; Wallace, D.; Williams, J.; Yamamoto, H.; Yvon-Lewis, S.; Yokouchi, Y.Atmospheric Chemistry and Physics (2013), 13 (17), 8915-8934, 20 pp.CODEN: ACPTCE; ISSN:1680-7324. (Copernicus Publications)Volatile halogenated org. compds. contg. bromine and iodine, which are naturally produced in the ocean, are involved in ozone depletion in both the troposphere and stratosphere. Three prominent compds. transporting large amts. of marine halogens into the atm. are bromoform (CHBr3), dibromomethane (CH2Br2) and Me iodide (CH3I). The input of marine halogens to the stratosphere has been estd. from observations and modeling studies using low-resoln. oceanic emission scenarios derived from top-down approaches. In order to improve emission inventory ests., we calc. data-based high resoln. global sea-to-air flux ests. of these compds. from surface observations within the HalOcAt (Halocarbons in the Ocean and Atm.) database. Global maps of marine and atm. surface concns. are derived from the data which are divided into coastal, shelf and open ocean regions. Considering phys. and biogeochem. characteristics of ocean and atm., the open ocean water and atm. data are classified into 21 regions. The available data are interpolated onto a 1° × 1° grid while missing grid values are interpolated with latitudinal and longitudinal dependent regression techniques reflecting the compds.' distributions. With the generated surface concn. climatologies for the ocean and atm., global sea-to-air concn. gradients and sea-to-air fluxes are calcd. Based on these calcns. we est. a total global flux of 1.5/2.5 Gmol Br yr-1 for CHBr3, 0.78/0.98 Gmol Br yr-1 for CH2Br2 and 1.24/1.45 Gmol Br yr-1 for CH3I (robust fit/ordinary least squares regression techniques). Contrary to recent studies, neg. fluxes occur in each sea-to-air flux climatol., mainly in the Arctic and Antarctic regions. "Hot spots" for global polybromomethane emissions are located in the equatorial region, whereas Me iodide emissions are enhanced in the subtropical gyre regions. Inter-annual and seasonal variation is contained within our flux calcns. for all three compds. Compared to earlier studies, our global fluxes are at the lower end of ests., esp. for bromoform. An under-representation of coastal emissions and of extreme events in our est. might explain the mismatch between our bottom-up emission est. and top-down approaches.
- 45Nightingale, P. D.; Malin, G.; Law, C. S.; Watson, A. J.; Liss, P. S.; Liddicoat, M. I.; Boutin, J.; Upstill-Goddard, R. C. In situ evaluation of air-sea gas exchange parameterizations using novel conservative and volatile tracers. Global Biogeochemical Cycles 2000, 14, 373– 387, DOI: 10.1029/1999GB900091Google Scholar45https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXhvVGms7s%253D&md5=d317f0ad2e899416471475672bd335f3In situ evaluation of air-sea gas exchange parameterizations using novel conservative and volatile tracersNightingale, Philip D.; Malin, Gill; Law, Cliff S.; Watson, Andrew J.; Liss, Peter S.; Liddicoat, Malcolm I.; Boutin, Jacqueline; Upstill-Goddard, Robert C.Global Biogeochemical Cycles (2000), 14 (1), 373-387CODEN: GBCYEP; ISSN:0886-6236. (American Geophysical Union)Measurements of air-sea gas exchange rates are reported from 2 deliberate tracer expts. in the southern North Sea during Feb. 1992 and 1993. A conservative tracer, spores of Bacillus globigii var. Niger, was used for the 1st time in an in situ air-sea gas exchange expt. This nonvolatile tracer is used to correct for dispersive diln. of the volatile tracers and allows 3 estns. of the transfer velocity for the same time period. The 1st estn. of the power dependence of gas transfer on mol. diffusivity in the marine environment is reported. This allows the impact of bubbles on ests. of the transfer velocity derived from changes in the He/SF6 ratio to be assessed. Data from earlier dual tracer expts. are reinterpreted, and findings suggest that results from all dual tracer expts. are mutually consistent. The complete data set is used to test published parameterizations of gas transfer with wind speed. A gas exchange relation that shows a dependence on wind speed intermediate between those of Liss and Merlivat (1986) and Wanninkhof (1992) is optimal. The dual tracer data are shown to be reasonably consistent with global ests. of gas exchange based on the uptake of natural and bomb-derived radiocarbon. The degree of scatter in the data when plotted against wind speed suggests that parameters not scaling with wind speed are also influencing gas exchange rates.
- 46Lana, A.; Bell, T. G.; Simó, R.; Vallina, S. M.; Ballabrera-Poy, J.; Kettle, A. J.; Dachs, J.; Bopp, L.; Saltzman, E. S.; Stefels, J. An updated climatology of surface dimethlysulfide concentrations and emission fluxes in the global ocean. Global Biogeochemical Cycles 2011, 25, 886, DOI: 10.1029/2010GB003850Google ScholarThere is no corresponding record for this reference.
- 47Granier, C. S.; Darras, H.; Denier van der Gon, J.; Doubalova, N.; Elguindi, B.; Galle, M.; Gauss, M.; Guevara, J.; Jalkanen, J.; Kuenen, C.; Liousse, B.; Quack, D.; Simpson, K. The Copernicus Atmosphere Monitoring Service global and regional emissions. Sindelarova The Copernicus Atmosphere Monitoring Service global and regional emissions (April 2019 version) 2019, 16, DOI: 10.24380/d0bn-kx16Google ScholarThere is no corresponding record for this reference.
- 48Sindelarova, K.; Granier, C.; Bouarar, I.; Guenther, A.; Tilmes, S.; Stavrakou, T.; Muller, J.-F.; Kuhn, U.; Stefani, P.; Knorr, W. Global dataset of biogenic VOC emissions calculated by the MEGAN model over the last 30 years. Atmospheric Chemistry and Physics 2014, 14, 9317, DOI: 10.5194/acp-14-9317-2014Google Scholar48https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhs1yktbjO&md5=978d8465b40240c42347d621604c1516Global data set of biogenic VOC emissions calculated by the MEGAN model over the last 30 yearsSindelarova, K.; Granier, C.; Bouarar, I.; Guenther, A.; Tilmes, S.; Stavrakou, T.; Muller, J.-F.; Kuhn, U.; Stefani, P.; Knorr, W.Atmospheric Chemistry and Physics (2014), 14 (17), 9317-9341, 25 pp.CODEN: ACPTCE; ISSN:1680-7324. (Copernicus Publications)The Model of Emissions of Gases and Aerosols from Nature (MEGANv2.1) together with the Modern-Era Retrospective Anal. for Research and Applications (MERRA) meteorol. fields were used to create a global emission data set of biogenic volatile org. compds. (BVOC) available on a monthly basis for the time period of 1980-2010. This data set, developed under the Monitoring Atm. Compn. and Climate project (MACC), is called MEGAN-MACC. The model estd. mean annual total BVOC emission of 760 Tg (C) yr-1 consisting of isoprene (70%), monoterpenes (11%), methanol (6%), acetone (3%), sesquiterpenes (2.5%) and other BVOC species each contributing less than 2%. Several sensitivity model runs were performed to study the impact of different model input and model settings on isoprene ests. and resulted in differences of up to ±17% of the ref. isoprene total. A greater impact was obsd. for a sensitivity run applying parameterization of soil moisture deficit that led to a 50% redn. of isoprene emissions on a global scale, most significantly in specific regions of Africa, South America and Australia. MEGAN-MACC ests. are comparable to results of previous studies. More detailed comparison with other isoprene inventories indicated significant spatial and temporal differences between the data sets esp. for Australia, Southeast Asia and South America. MEGAN-MACC ests. of isoprene, α-pinene and group of monoterpenes showed a reasonable agreement with surface flux measurements at sites located in tropical forests in the Amazon and Malaysia. The model was able to capture the seasonal variation of isoprene emissions in the Amazon forest.
- 49Smith, B.; Wårlind, D.; Arneth, A.; Hickler, T.; Leadley, P.; Siltberg, J.; Zaehle, S. Implications of incorporating N cycling and N limitations on primary production in an individual-based dynamic vegetation model. Biogeosciences 2014, 11, 2027– 2054, DOI: 10.5194/bg-11-2027-2014Google ScholarThere is no corresponding record for this reference.
- 50Arneth, A.; Niinemets, U.; Pressley, S.; Bäck, J.; Hari, P.; Karl, T.; Noe, S.; Prentice, I. C.; Serça, D.; Hickler, T.; Wolf, A.; Smith, B. Process-based estimates of terrestrial ecosystem isoprene emissions: incorporating the effects of a direct CO2-isoprene interaction. Atmospheric Chemistry and Physics 2007, 7, 31– 53, DOI: 10.5194/acp-7-31-2007Google ScholarThere is no corresponding record for this reference.
- 51Paulot, F.; Jacob, D. J.; Johnson, M. T.; Bell, T. G.; Baker, A. R.; Keene, W. C.; Lima, I. D.; Doney, S. C.; Stock, C. A. Global oceanic emission of ammonia: Constraints from seawater and atmospheric observations. Global Biogeochemical Cycles 2015, 29, 1165– 1178, DOI: 10.1002/2015GB005106Google ScholarThere is no corresponding record for this reference.
- 52Carpenter, L.; MacDonald, S.; Shaw, M.; Kumar, R.; Saunders, R.; Parthipan, R.; Wilson, J.; Plane, J. Atmospheric iodine levels influenced by sea surface emissions of inorganic iodine. Nature Geoscience 2013, 6, 108– 111, DOI: 10.1038/ngeo1687Google Scholar52https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXntFyqsw%253D%253D&md5=fccd3f2ce8410029ba975238eb11da97Atmospheric iodine levels influenced by sea surface emissions of inorganic iodineCarpenter, Lucy J.; MacDonald, Samantha M.; Shaw, Marvin D.; Kumar, Ravi; Saunders, Russell W.; Parthipan, Rajendran; Wilson, Julie; Plane, John M. C.Nature Geoscience (2013), 6 (2), 108-111CODEN: NGAEBU; ISSN:1752-0894. (Nature Publishing Group)Naturally occurring bromine- and iodine-contg. compds. substantially reduce regional, and possibly even global, tropospheric ozone levels. As such, these halogen gases reduce the global warming effects of ozone in the troposphere, and its capacity to initiate the chem. removal of hydrocarbons such as methane. The majority of halogen-related surface ozone destruction is attributable to iodine chem. So far, org. iodine compds. have been assumed to serve as the main source of oceanic iodine emissions. However, known org. sources of atm. iodine cannot account for gas-phase iodine oxide concns. in the lower troposphere over the tropical oceans. Here, we quantify gaseous emissions of inorg. iodine following the reaction of iodide with ozone in a series of lab. expts. We show that the reaction of iodide with ozone leads to the formation of both mol. iodine and hypoiodous acid. Using a kinetic box model of the sea surface layer and a one-dimensional model of the marine boundary layer, we show that the reaction of ozone with iodide on the sea surface could account for around 75% of obsd. iodine oxide levels over the tropical Atlantic Ocean. According to the sea surface model, hypoiodous acid-not previously considered as an oceanic source of iodine-is emitted at a rate ten-fold higher than that of mol. iodine under ambient conditions.
- 53Sofiev, M.; Soares, J.; Prank, M.; de Leeuw, G.; Kukkonen, J. A regional-to-global model of emission and transport of sea salt particles in the atmosphere. Journal of Geophysical Research (Atmospheres) 2011, 116, 21302, DOI: 10.1029/2010JD014713Google ScholarThere is no corresponding record for this reference.
- 54Gantt, B.; Meskhidze, N.; Facchini, M. C.; Rinaldi, M.; Ceburnis, D.; O'Dowd, C. D. Wind speed dependent size-resolved parameterization for the organic mass fraction of sea spray aerosol. Atmospheric Chemistry and Physics - ATMOS CHEM PHYS 2011, 11, 8777– 8790, DOI: 10.5194/acp-11-8777-2011Google ScholarThere is no corresponding record for this reference.
- 55Olenius, T.; Roldin, P. Role of gas-molecular cluster-aerosol dynamics in atmospheric new-particle formation. Sci. Rep. 2022, 12, 10135, DOI: 10.1038/s41598-022-14525-yGoogle ScholarThere is no corresponding record for this reference.
- 56Ning, A.; Liu, L.; Zhang, S.; Yu, F.; Du, L.; Ge, M.; Zhang, X. The critical role of dimethylamine in the rapid formation of iodic acid particles in marine areas. npj Climate and Atmospheric Science 2022, 5, 92, DOI: 10.1038/s41612-022-00316-9Google ScholarThere is no corresponding record for this reference.
- 57Besel, V.; Kubecka, J.; Kurtén, T.; Vehkamäki, H. Impact of Quantum Chemistry Parameter Choices and Cluster Distribution Model Settings on Modeled Atmospheric Particle Formation Rates. J. Phys. Chem. A 2020, 124, 5931, DOI: 10.1021/acs.jpca.0c03984Google Scholar57https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXht1alsLvK&md5=8e749fba9da93b18694e415c1c060662Impact of Quantum Chemistry Parameter Choices and Cluster Distribution Model Settings on Modeled Atmospheric Particle Formation RatesBesel, Vitus; Kubecka, Jakub; Kurten, Theo; Vehkamaki, HannaJournal of Physical Chemistry A (2020), 124 (28), 5931-5943CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Various parameters effect on the rate of new particle formation predicted for a H2SO4-NH3 system using quantum chem. and cluster distribution dynamics simulations, here, Atm. Cluster Dynamics Code, was tested. Consistent consideration of the rotational symmetry no. of monomers (H2SO4 and NH3 mols., bisulfate ion and NH4+) led to a significant rise in predicted particle formation rate; including the rotational symmetry no. of clusters only slightly changed results, and only for conditions where charged clusters dominate particle formation rate. This was because most of clusters stable enough to participate in new particle formation have a rotational symmetry no. of 1; few exceptions to this rule are pos. charged clusters. Applying quasi-harmonic correction for low frequency vibrational modes tended to generally decrease predicted new particle formation rates and significantly altered the slope of the formation rate curve plotted against H2SO4 concn., a typical convention in atm. aerosol science. Cluster max. size effect explicitly included in simulations depended on simulation conditions. Errors arising from a limited set of clusters were higher for higher evapn. rates and tended increase with temp. Errors tended to be higher for lower vapor concns. Boundary conditions for out-growing clusters (counted as formed particles) had only a small effect on results, provided the definition was chem. reasonable and the set of simulated clusters was sufficiently large. A comparison of data from cosmics leaving outdoor droplets (CLOUD) chamber and a cluster distribution dynamics model using older quantum chem. input data showed improved agreement when using new input data and the proposed combination of symmetry/quasi-harmonic corrections.
- 58Myllys, N.; Kubečka, J.; Besel, V.; Alfaouri, D.; Olenius, T.; Smith, J. N.; Passananti, M. Role of base strength, cluster structure and charge in sulfuric-acid-driven particle formation. Atmospheric Chemistry and Physics 2019, 19, 9753– 9768, DOI: 10.5194/acp-19-9753-2019Google ScholarThere is no corresponding record for this reference.
- 59Zhang, R.; Xie, H.-B.; Ma, F.; Chen, J.; Iyer, S.; Simon, M.; Heinritzi, M.; Shen, J.; Tham, Y. J.; Kurtén, T.; Worsnop, D. R.; Kirkby, J.; Curtius, J.; Sipilä, M.; Kulmala, M.; He, X.-C. Critical Role of Iodous Acid in Neutral Iodine Oxoacid Nucleation. Environ. Sci. Technol. 2022, 56, 14166– 14177, DOI: 10.1021/acs.est.2c04328Google Scholar59https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB38XisVajsLrM&md5=529b05e9e166104af0dc3f36a5a4445fCritical Role of Iodous Acid in Neutral Iodine Oxoacid NucleationZhang, Rongjie; Xie, Hong-Bin; Ma, Fangfang; Chen, Jingwen; Iyer, Siddharth; Simon, Mario; Heinritzi, Martin; Shen, Jiali; Tham, Yee Jun; Kurten, Theo; Worsnop, Douglas R.; Kirkby, Jasper; Curtius, Joachim; Sipila, Mikko; Kulmala, Markku; He, Xu-ChengEnvironmental Science & Technology (2022), 56 (19), 14166-14177CODEN: ESTHAG; ISSN:1520-5851. (American Chemical Society)Nucleation of neutral iodine particles has recently been found to involve both iodic acid (HIO3) and iodous acid (HIO2). However, the precise role of HIO2 in iodine oxoacid nucleation remains unclear. Herein, we probe such a role by investigating the cluster formation mechanisms and kinetics of (HIO3)m(HIO2)n (m = 0-4, n = 0-4) clusters with quantum chem. calcns. and atm. cluster dynamics modeling. When compared with HIO3, we find that HIO2 binds more strongly with HIO3 and also more strongly with HIO2. After accounting for ambient vapor concns., the fastest nucleation rate is predicted for mixed HIO3-HIO2 clusters rather than for pure HIO3 or HIO2 ones. Our calcns. reveal that the strong binding results from HIO2 exhibiting a base behavior (accepting a proton from HIO3) and forming stronger halogen bonds. Moreover, the binding energies of (HIO3)m(HIO2)n clusters show a far more tolerant choice of growth paths when compared with the strict stoichiometry required for sulfuric acid-base nucleation. Our predicted cluster formation rates and dimer concns. are acceptably consistent with those measured by the Cosmic Leaving Outdoor Droplets (CLOUD) expt. This study suggests that HIO2 could facilitate the nucleation of other acids beyond HIO3 in regions where base vapors such as ammonia or amines are scarce.
- 60Schmitz, G.; Elm, J. Assessment of the DLPNO Binding Energies of Strongly Noncovalent Bonded Atmospheric Molecular Clusters. ACS Omega 2020, 5, 7601– 7612, DOI: 10.1021/acsomega.0c00436Google Scholar60https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXlt1yktL0%253D&md5=865ba10b3e9832494d960adb64b36715Assessment of the DLPNO Binding Energies of Strongly Noncovalent Bonded Atmospheric Molecular ClustersSchmitz, Gunnar; Elm, JonasACS Omega (2020), 5 (13), 7601-7612CODEN: ACSODF; ISSN:2470-1343. (American Chemical Society)This work assessed the performance of DLPNO-CCSD(T0), DLPNO-MP2, and d. functional theory methods in calcg. binding energies of a representative test set of 45 atm. acid/acid, acid/base, and acid/water dimer clusters. Approx. method performance was compared to high level, explicitly correlated CCSD(F12*)(T)/complete basis set (CBS) ref. calcns. Of the tested d. functionals, ωB97X-D3(BJ) had the best performance with a mean deviation of 0.09 kcal/mol and a max. deviation of 0.83 kcal/mol. The RI-CC2/aug-cc-pV(T+d)Z level of theory severely over-predicted cluster binding energies with a mean deviation of -1.31 kcal/mol and a max. deviation up to -3.00 kcal/mol. Thus, RI-CC2/aug-cc-pV(T+d)Z should not be used to study atm. mol. clusters. DLPNO variants were tested with/without the inclusion of explicit correlation (F12) in the wave function, with different pair natural orbital (PNO) settings (loosePNO, normalPNO, tightPNO) and using double and triple zeta basis sets. Performance of DLPNO-MP2 methods was independent of PNO settings and yielded low mean deviations (≤-0.84 kcal/mol). However, DLPNO-MP2 required explicitly correlated wave functions to yield max. deviations <1.40 kcal/mol. To obtain high accuracy with max. deviation <∼1.0 kcal/mol, DLPNO-CCSD(T0)/aug-cc-pVTZ (normalPNO) calcns. or DLPNO-CCSD(T0)-F12/cc-pVTZ-F12 (normalPNO) calcns. were required. The most accurate level of theory was DLPNO-CCSD(T0)-F12/cc-pVTZ-F12 using a tightPNO criterion which yielded a mean deviation of 0.10 kcal/mol, with a max. deviation of 0.20 kcal/mol, vs. the CCSD(F12*)(T)/CBS ref.
- 61Elm, J.; Kristensen, K. Basis Set Convergence of the Binding Energies of Strongly Hydrogen-Bonded Atmospheric Clusters. Phys. Chem. Chem. Phys. 2017, 19, 1122– 1133, DOI: 10.1039/C6CP06851KGoogle Scholar61https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XhvFKjtb%252FO&md5=c44dbe667e0a1605a94c77283fbf5b08Basis set convergence of the binding energies of strongly hydrogen-bonded atmospheric clustersElm, Jonas; Kristensen, KasperPhysical Chemistry Chemical Physics (2017), 19 (2), 1122-1133CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)We investigate the basis set convergence of second order Moller Plesset perturbation theory for the binding energies of 11 strongly hydrogen-bonded clusters relevant to the atm. The binding energies are calcd. both as the uncorrected (UC) value, as well as by employing the counterpoise (CP) and the Same No. Of Optimized Parameters (SNOOP) correction schemes, with and without explicit correlation (F12). We find that the use of the F12 corrections is of utter importance for obtaining converged binding energies. Without F12 corrections at least a quadruple-ζ basis set is generally required to yield errors below 1 kcal mol-1 compared to the complete basis set (CBS) limit. Using coupled cluster methods we obtain a best est. of the CCSD(T) CBS limit of the binding energies of the considered clusters and compare it with approx. DLPNO-CCSD(T) and DFT methods. We identify a pragmatic approach, relying solely on a series of double-ζ basis set calcns., for obtaining results in remarkably good agreement with our CBS est.
- 62Klamt, A. Conductor-like screening model for real solvents: a new approach to the quantitative calculation of solvation phenomena. J. Phys. Chem. 1995, 99, 2224– 2235, DOI: 10.1021/j100007a062Google Scholar62https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXjsFaisb0%253D&md5=74ff536ca20d353b9f8525d9f25c51d5Conductor-like Screening Model for Real Solvents: A New Approach to the Quantitative Calculation of Solvation PhenomenaKlamt, AndreasJournal of Physical Chemistry (1995), 99 (7), 2224-35CODEN: JPCHAX; ISSN:0022-3654. (American Chemical Society)Starting from the question of why dielec. continuum models give a fairly good description of mols. in water and some other solvents, a totally new approach for the calcn. of solvation phenomena is presented. It is based on the perfect, i.e., conductor-like, screening of the solute mol. and a quant. calcn. of the deviations from ideality appearing in real solvents. Thus, a new point of view to solvation phenomena is presented, which provides an alternative access to many questions of scientific and tech. importance. The whole theory is based on the results of MO continuum solvation models. A few representative solvents are considered, and the use of the theory is demonstrated by the calcn. of vapor pressures, surface tensions, and octanol/water partition coeffs.
- 63Klamt, A.; Jonas, V.; Bürger, T.; Lohrenz, J. C. W. Refinement and parametrization of COSMO-RS. J. Phys. Chem. A 1998, 102, 5074– 5085, DOI: 10.1021/jp980017sGoogle Scholar63https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXjs1Srs7Y%253D&md5=e9043dfbc9123e2a9b7e7a77216db911Refinement and Parametrization of COSMO-RSKlamt, Andreas; Jonas, Volker; Buerger, Thorsten; Lohrenz, John C. W.Journal of Physical Chemistry A (1998), 102 (26), 5074-5085CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The continuum solvation model COSMO and its extension beyond the dielec. approxn. (COSMO-RS) have been carefully parametrized in order to optimally reproduce 642 data points for a variety of properties, i.e., ΔG of hydration, vapor pressure, and the partition coeffs. for octanol/water, benzene/water, hexane/water, and di-Et ether/water. Two hundred seventeen small to medium sized neutral mols., covering most of the chem. functionality of the elements H, C, N, O, and Cl, have been considered. An overall accuracy of 0.4 (rms) kcal/mol for chem. potential differences, corresponding to a factor of 2 in the equil. consts. under consideration, has been achieved. This was using only a single radius and one dispersion const. per element and a total no. of eight COSMO-RS inherent parameters. Most of these parameters were close to their theor. est. The optimized cavity radii agreed well with the widely accepted rule of 120% of van der Waals radii. The whole parametrization was based upon d. functional calcns. using DMol/COSMO. As a result of this sound parametrization, we are now able to calc. almost any chem. equil. in liq./liq. and vapor/liq. systems up to an accuracy of a factor 2 without the need of any addnl. exptl. data for solutes or solvents. This opens a wide range of applications in phys. chem. and chem. engineering.
- 64Eckert, F.; Klamt, A. Fast solvent screening via quantum chemistry: COSMO-RS approach. AIChE J. 2002, 48, 369– 385, DOI: 10.1002/aic.690480220Google Scholar64https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD38XhsFSksLg%253D&md5=0f1996a34219de62fa735128a95020a1Fast solvent screening via quantum chemistry: COSMO-RS approachEckert, Frank; Klamt, AndreasAIChE Journal (2002), 48 (2), 369-385CODEN: AICEAC; ISSN:0001-1541. (American Institute of Chemical Engineers)Conductor-like Screening Model for Real Solvents (COSMO-RS), a general and fast methodol. for the a priori prediction of thermophys. data of liqs. is presented. It is based on cheap unimol. quantum chem. calcns., which, combined with exact statistical thermodn., provide the information necessary for the evaluation of mol. interactions in liqs. COSMO-RS is an alternative to structure interpolating group contribution methods. The method is independent of exptl. data and generally applicable. A methodol. comparison with group contribution methods is given. The applicability of the COSMO-RS method to the goal of solvent screening is demonstrated at various examples of vapor-liq.-, liq.-liq.-, solid-liq.-equil. and vapor-pressure predictions.
- 65BIOVIA COSMOtherm, Release 2021; Dassault Systèmes. http://www.3ds.com. 2021.Google ScholarThere is no corresponding record for this reference.
- 66BIOVIA COSMOconf 2021; Dassault Systèmes. http://www.3ds.com. 2021.Google ScholarThere is no corresponding record for this reference.
- 67Turbomole, V7.5.1; University of Karlsruhe and Forschungszentrum Karlsruhe GmbH, 2020.Google ScholarThere is no corresponding record for this reference.
- 68Roldin, P.; Swietlicki, E.; Massling, A.; Kristensson, A.; Löndahl, J.; Eriksson, A.; Pagels, J.; Gustafsson, S. Aerosol ageing in an urban plume - implication for climate. Atmospheric Chemistry and Physics - ATMOS CHEM PHYS 2011, 11, 5897– 5915, DOI: 10.5194/acp-11-5897-2011Google ScholarThere is no corresponding record for this reference.
- 69Jacobson, M. Z. Fundamentals of Atmospheric Modeling, 2nd ed.; Cambridge University Press, 2005.Google ScholarThere is no corresponding record for this reference.
- 70BIOVIA COSMOtherm, version C3.0, Release 19; Dassault Systemes, 2019.Google ScholarThere is no corresponding record for this reference.
- 71Pruppacher, H.; Klett, J.; Wang, P. Microphysics of Clouds and Precipitation. Aerosol Sci. Technol. 1998, 28, 381– 382, DOI: 10.1080/02786829808965531Google ScholarThere is no corresponding record for this reference.
- 72Kurosaki, Y.; Matoba, S.; Iizuka, Y.; Fujita, K.; Shimada, R. Increased oceanic dimethyl sulfide emissions in areas of sea ice retreat inferred from a Greenland ice core. Communications Earth & Environment 2022, 3, 327, DOI: 10.1038/s43247-022-00661-wGoogle ScholarThere is no corresponding record for this reference.
Cited By
Smart citations by scite.ai include citation statements extracted from the full text of the citing article. The number of the statements may be higher than the number of citations provided by ACS Publications if one paper cites another multiple times or lower if scite has not yet processed some of the citing articles.
This article is cited by 4 publications.
- Jiangyi Zhang, Jian Zhao, Robin Wollesen de Jonge, Nina Sarnela, Pontus Roldin, Mikael Ehn. Evaluating the Applicability of a Real-Time Highly Oxygenated Organic Molecule (HOM)-Based Indicator for Ozone Formation Sensitivity at a Boreal Forest Station. Environmental Science & Technology Letters 2024, 11
(11)
, 1227-1232. https://doi.org/10.1021/acs.estlett.4c00733
- Kuanyun Hu, Jie Hu, Narcisse Tsona Tchinda, Christian George, Jianlong Li, Lin Du. Revealing the composition and optical properties of marine carbonaceous aerosols: A case of the eastern China marginal seas. Science of The Total Environment 2025, 958 , 178136. https://doi.org/10.1016/j.scitotenv.2024.178136
- Xueqi Ma, Kun Li, Shan Zhang, Narcisse Tsona Tchinda, Jianlong Li, Hartmut Herrmann, Lin Du. Molecular characteristics of sea spray aerosols during aging with the participation of marine volatile organic compounds. Science of The Total Environment 2024, 954 , 176380. https://doi.org/10.1016/j.scitotenv.2024.176380
- Yu Jiang, Juan Yu, Ji-Yuan Tian, Gui-Peng Yang, Long-Fei Liu, Xin-Ran Song, Rong Chen. Microplastics and copper impacts on feeding, oxidative stress, antioxidant enzyme activity, and dimethylated sulfur compounds production in Manila clam Ruditapes philippinarum. Marine Pollution Bulletin 2024, 208 , 117022. https://doi.org/10.1016/j.marpolbul.2024.117022
Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.
Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.
The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.
Recommended Articles
Abstract
Figure 1
Figure 1. Effect of marine versus continental air mass history on the aerosol number size distribution in a boreal forest environment. Panel (a) depicts the typical summertime source regions of DMS (purple), MTs (green), NH3 (blue), and anthropogenic SO2 (yellow), along with their position in relation to the Hyltemossa, SMEARII, and Pallas measurements stations. These colors indicate only the spatial locations of the emissions of the different species between May and August, 2018. Panels (b), (c), and (d) show the spring and summertime (March-August) mean aerosol number size distributions measured at the Hyltemossa, SMEARII, and Pallas stations during the periods 2018–2020, 2006–2020, and 2015–2018, respectively. The size distributions in panel (b), (c,) and (d) are separated based on the marine influence on their air mass history.
Figure 2
Figure 2. Modeled and measured long-term and diurnal gas-phase concentrations of MTs, HOM monomers, isoprene, SA, MSA, and O3 at the Station for Measuring Ecosystem-Atmosphere Relations II (SMEARII) between the 17th of May and 28th of August, 2018. Gray and black dots denote the surface and above-canopy measurements, respectively, while the colored lines represent the above-canopy model results. Marine periods 1, 2, and 3 (MP1, MP2, and MP3) denote periods of particularly high marine air-mass impact, while continental period 1 (CP1) denotes a period of particularly high continental air-mass impact. The back-trajectory heat-maps displayed above each period illustrate the regions from which the air-masses arrived during MP1, MP2, MP3, and CP1. The shaded areas denote the measured and modeled data range within the 25th and 75th percentile.
Figure 3
Figure 3. Measured and modeled (a–e) time-dependent and (f–o) median particle number size distributions at the Station for Measuring Ecosystem Atmosphere Relations II (SMEARII) between the 17th of May and 28th of August. Marine periods 1, 2, and 3 (MP1, MP2, and MP3) denote periods of particular high marine air-mass impact, while continental period 1 (CP1) denotes a period of particular high continental air-mass impact. The back-trajectory heat-maps displayed above each period illustrates the regions from which the air-masses arrived during MP1, MP2, MP3, and CP1. The median size distributions are separated into (f–j) periods of predominant marine air-mass impact (time spent over the ocean >50th prct., purple), and (k–o) periods of predominant continental impact (time spent over the ocean <50th prct., green). The averaging for the median size distributions is done for the whole simulation period and not just during MP1, MP2, MP3, and CP1. The model results include data from the base case run (BaseCase), the without DMS emissions simulation (woDMS), the without iodine nucleation simulation (woIodine), and the without anthropogenic emissions simulation (woAnthro). The shaded areas denote the measured and modeled data range within the 25th and 75th percentile.
Figure 4
Figure 4. Schematic depiction of the influence of DMS and various iodine species on the formation and growth of aerosol particles both over the ocean and over land. (1) DMS, CH3I, I2, and HOI oxidize over the ocean and land to form the strong acids SA, MSA, HIO3, and HIO2, which nucleate (3) among themselves or (2) in the presence of NH3 or DMA. These particles in turn are (4) grown by condensation of low volatile organic compounds over the boreal forest, ultimately reaching (5) the CCN size range, where they (6) affect the formation, lifetime, and precipitation of clouds. Created with BioRender.com.
References
This article references 72 other publications.
- 1Lovelock, J. E.; Maggs, R. J.; Rasmussen, R. A. Atmospheric Dimethyl Sulfide and the Natural Sulfur Cycle. Nature 1972, 237, 452– 453, DOI: 10.1038/237452a01https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE3sXisVOmug%253D%253D&md5=c6d0495b981f24e9fef6102dd86617eeAtmospheric dimethyl sulfide and the natural sulfur cycleLovelock, J. E.; Maggs, R. J.; Rasmussen, R. A.Nature (London, United Kingdom) (1972), 237 (5356), 452-3CODEN: NATUAS; ISSN:0028-0836.The av. concn. of dimethyl sulfide (DMS) [75-18-3] in seawater was 1.2 .tim. 10-11 g/ml and the distribution coeff. of DMS between air and seawater was 0.30, corresponding to an atm. concn. of 1.2 ppb at equil. The emission rates of DMS were 50-400 .tim. 10-12 g/g wet wt./hr from marine algae, 21-84 .tim. 10-12 g/g/hr from soils, and 2-43 .tim. 10-12 g/g dry wt./hr from the living intact leaves of oak, cotton, spruce, and pine trees. DMS may be involved in the transfer of biol. sulfur [7704-34-9] in nature.
- 2Carpenter, L.; Archer, S.; Beale, R. Ocean-atmosphere trace gas exchange. Chem. Soc. Rev. 2012, 41, 6473– 506, DOI: 10.1039/c2cs35121h2https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xhtlaktr%252FP&md5=12993bba4892e6ceadc7b6dc4d8e9432Ocean-atmosphere trace gas exchangeCarpenter, Lucy J.; Archer, Stephen D.; Beale, RachaelChemical Society Reviews (2012), 41 (19), 6473-6506CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. The oceans contribute significantly to the global emissions of a no. of atmospherically important volatile gases, notably those contg. sulfur, nitrogen, and halogens. Such gases play crit. roles not only in global biogeochem. cycling but also in a wide range of atm. processes including marine aerosol formation and modification, tropospheric ozone formation and destruction, photooxidant cycling, and stratospheric ozone loss. A no. of marine emissions are greenhouse gases, others influence the Earth's radiative budget indirectly through aerosol formation and/or by modifying oxidant levels and thus changing the atm. lifetime of gases such as methane. Reviewed is current literature concerning the phys., chem., and biol. controls on the sea-air emissions of a wide range of gases including di-Me sulfide (DMS), halocarbons, nitrogen-contg. gases including NH3, amines (including dimethylamine, DMA, and diethylamine, DEA), alkyl nitrates (RONO2) and N2O, non-methane hydrocarbons (NMHC) including isoprene and oxygenated (O)VOCs, CH4 and CO. Where possible the current global emission budgets of these gases as well as known mechanisms for their formation and loss in the surface ocean are reviewed.
- 3Zheng, G.; Wang, Y.; Wood, R.; Jensen, M. P.; Kuang, C.; McCoy, I. L.; Matthews, A.; Mei, F.; Tomlinson, J. M.; Shilling, J. E. New particle formation in the remote marine boundary layer. Nat. Commun. 2021, 12, 527, DOI: 10.1038/s41467-020-20773-13https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXit1alsb4%253D&md5=cb0000d7b4413432690fd0ddf779d694New particle formation in the remote marine boundary layerZheng, Guangjie; Wang, Yang; Wood, Robert; Jensen, Michael P.; Kuang, Chongai; McCoy, Isabel L.; Matthews, Alyssa; Mei, Fan; Tomlinson, Jason M.; Shilling, John E.; Zawadowicz, Maria A.; Crosbie, Ewan; Moore, Richard; Ziemba, Luke; Andreae, Meinrat O.; Wang, JianNature Communications (2021), 12 (1), 527CODEN: NCAOBW; ISSN:2041-1723. (Nature Research)Abstr.: Marine low clouds play an important role in the climate system, and their properties are sensitive to cloud condensation nuclei concns. While new particle formation represents a major source of cloud condensation nuclei globally, the prevailing view is that new particle formation rarely occurs in remote marine boundary layer over open oceans. Here we present evidence of the regular and frequent occurrence of new particle formation in the upper part of remote marine boundary layer following cold front passages. The new particle formation is facilitated by a combination of efficient removal of existing particles by pptn., cold air temps., vertical transport of reactive gases from the ocean surface, and high actinic fluxes in a broken cloud field. The newly formed particles subsequently grow and contribute substantially to cloud condensation nuclei in the remote marine boundary layer and thereby impact marine low clouds.
- 4Barnes, I.; Hjorth, J.; Mihalopoulos, N. Dimethyl Sulfide and Dimethyl Sulfoxide and Their Oxidation in the Atmosphere. Chem. Rev. 2006, 106, 940– 975, DOI: 10.1021/cr020529+4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28Xhs1Grtr4%253D&md5=c2a51fd2de3a72676c30eb83aeed4fb1Dimethyl Sulfide and Dimethyl Sulfoxide and Their Oxidation in the AtmosphereBarnes, Ian; Hjorth, Jens; Mihalopoulos, NikosChemical Reviews (Washington, DC, United States) (2006), 106 (3), 940-975CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review concerning the gas phase chem., atm. oxidn., and chem. products of the atm. oxidn. of di-Me sulfide (DMS) and dimethylsulfoxide (DMSO) is given. Topics discussed include: DMS chem. (reaction with OH-, NO3-, halogen atoms and halogen oxides [kinetics, primary reaction mechanisms, reaction products]); DMSO chem. (reaction with OH-, NO3-, halogen atoms and halogen oxides [kinetics, primary reaction mechanisms, reaction products]); and current state of field measurements and their interpretation (multi-phase chem. involved in atm. oxidn. of DMS; field and modeling study evidence of the role of multi-phase reactions in the DMS cycle; aq.-phase reactions of DMS, DMSO, di-Me sulfone, methane sulfinic acid [MSIA], and methane sulfonic acid [MSA]; atm. implication of aq.-phase reactions of DMS, DMSO, MSIA, and MSA; and modeling study recommendations).
- 5Hoffmann, E. H.; Tilgner, A.; Schrödner, R.; Bräuer, P.; Wolke, R.; Herrmann, H. An advanced modeling study on the impacts and atmospheric implications of multiphase dimethyl sulfide chemistry. Proc. Natl. Acad. Sci. U. S. A. 2016, 113, 11776– 11781, DOI: 10.1073/pnas.16063201135https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XhsFyrsb%252FO&md5=97c54638e0a27847d0f959788895cd49An advanced modeling study on the impacts and atmospheric implications of multiphase dimethyl sulfide chemistryHoffmann, Erik Hans; Tilgner, Andreas; Schroedner, Roland; Braeuer, Peter; Wolke, Ralf; Herrmann, HartmutProceedings of the National Academy of Sciences of the United States of America (2016), 113 (42), 11776-11781CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)Oceans dominate di-Me sulfide (DMS) emissions, the major natural S source. DMS is important for formation of non-sea salt sulfate (nss-SO42-) aerosols and secondary particulate matter over oceans; thus, it significantly affects global climate. The DMS oxidn. mechanism has been examd. in several different model studies; however, these studies had restricted oxidn. mechanisms which mainly under-represented important aq.-phase chem. processes. These neglected but highly effective processes strongly affect the direct product yields of DMS oxidn., thereby affecting the climatic influence of aerosols. To address these shortfalls, an extensive multiphase DMS chem. mechanism, the Chem. Aq. Phase Radical Mechanism DMS Module 1.0, was developed and used in detailed model assessments of multiphase DMS chem. in the marine boundary layer. Model studies confirmed the importance of aq. phase chem. for the fate of DMS and its oxidn. products. Aq. phase processes significantly reduced the SO2 yield and increased the Me sulfonic acid (MSA) yield, which is needed to close the gap between modeled and measured MSA concns. Simulations implied that multi-phase DMS oxidn. produces equal amts. of MSA and SO42-, a result with significant implications for nss-SO42- aerosol formation, cloud condensation nuclei concn., and cloud albedo over oceans. Results showed the deficiencies of parameterizations currently used in higher-scale models which only treat gas phase chem. Overall, the results showed treatment of DMS chem. in gas and aq. phases is essential to improve model prediction accuracy.
- 6Wollesen de Jonge, R.; Elm, J.; Rosati, B.; Christiansen, S.; Hyttinen, N.; Lüdemann, D.; Bilde, M.; Roldin, P. Secondary aerosol formation from dimethyl sulfide - improved mechanistic understanding based on smog chamber experiments and modelling. Atmospheric Chemistry and Physics 2021, 21, 9955– 9976, DOI: 10.5194/acp-21-9955-2021There is no corresponding record for this reference.
- 7Gomez Martin, J. C.; Lewis, T. R.; Blitz, M. A.; Plane, J. M. C.; Kumar, M.; Francisco, J. S.; Saiz-Lopez, A. A gas-to-particle conversion mechanism helps to explain atmospheric particle formation through clustering of iodine oxides. Nat. Commun. 2020, 11, 4521, DOI: 10.1038/s41467-020-18252-8There is no corresponding record for this reference.
- 8Finkenzeller, H.; Iyer, S.; He, X.-C.; Simon, M.; Koenig, T. K.; Lee, C. F.; Valiev, R.; Hofbauer, V.; Amorim, A.; Baalbaki, R. The gas-phase formation mechanism of iodic acid as an atmospheric aerosol source. Nat. Chem. 2023, 15, 129– 135, DOI: 10.1038/s41557-022-01067-zThere is no corresponding record for this reference.
- 9Korhonen, P.; Kulmala, M.; Laaksonen, A.; Viisanen, Y.; McGraw, R.; Seinfeld, J. H. Ternary nucleation of H2SO4, NH3, and H2O in the atmosphere. Journal of Geophysical Research: Atmospheres 1999, 104, 26349– 26353, DOI: 10.1029/1999JD9007849https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXotFWqurk%253D&md5=d5ca8b589881057516ae0bb3ec62b4a9Ternary nucleation of H2SO4, NH3, and H2O in the atmosphereKorhonen, P.; Kulmala, M.; Laaksonen, A.; Viisanen, Y.; McGraw, R.; Seinfeld, J. H.Journal of Geophysical Research, [Atmospheres] (1999), 104 (D21), 26349-26353CODEN: JGRDE3 ISSN:. (American Geophysical Union)Classical theory of binary homogeneous nucleation is extended to the ternary system H2SO4-NH3-H2O. For NH3 mixing ratios exceeding about 1 ppt, the presence of NH3 enhances the binary H2SO4-H2O nucleation rate by several orders of magnitude. The Gibbs free energies of formation of the crit. H2SO4-NH3-H2O cluster, as calcd. by two independent approaches, are in substantial agreement. The finding that the H2SO4-NH3-H2O ternary nucleation rate is independent of relative humidity over a large range of H2SO4 concns. has wide atm. consequences. The limiting component for ternary H2SO4-NH3-H2O nucleation is, as in the binary H2SO4-H2O case, H2SO4; however, the H2SO4 concn. needed to achieve significant nucleation rates is several orders of magnitude below that required in the binary case.
- 10Almeida, J.; Schobesberger, S.; Kürten, A.; Ortega, I.; Kupiainen-Määttä, O.; Praplan, A.; Adamov, A.; Amorim, A.; Bianchi, F.; Breitenlechner, M. Molecular understanding of sulphuric acid-amine particle nucleation in the atmosphere. Nature 2013, 502, 359– 363, DOI: 10.1038/nature1266310https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhsFOit73N&md5=06460967339b7bcd3d231102fd2d96dbMolecular understanding of sulphuric acid-amine particle nucleation in the atmosphereAlmeida, Joao; Schobesberger, Siegfried; Kuerten, Andreas; Ortega, Ismael K.; Kupiainen-Maeaettae, Oona; Praplan, Arnaud P.; Adamov, Alexey; Amorim, Antonio; Bianchi, Federico; Breitenlechner, Martin; David, Andre; Dommen, Josef; Donahue, Neil M.; Downard, Andrew; Dunne, Eimear; Duplissy, Jonathan; Ehrhart, Sebastian; Flagan, Richard C.; Franchin, Alessandro; Guida, Roberto; Hakala, Jani; Hansel, Armin; Heinritzi, Martin; Henschel, Henning; Jokinen, Tuija; Junninen, Heikki; Kajos, Maija; Kangasluoma, Juha; Keskinen, Helmi; Kupc, Agnieszka; Kurten, Theo; Kvashin, Alexander N.; Laaksonen, Ari; Lehtipalo, Katrianne; Leiminger, Markus; Leppae, Johannes; Loukonen, Ville; Makhmutov, Vladimir; Mathot, Serge; McGrath, Matthew J.; Nieminen, Tuomo; Olenius, Tinja; Onnela, Antti; Petaejae, Tuukka; Riccobono, Francesco; Riipinen, Ilona; Rissanen, Matti; Rondo, Linda; Ruuskanen, Taina; Santos, Filipe D.; Sarnela, Nina; Schallhart, Simon; Schnitzhofer, Ralf; Seinfeld, John H.; Simon, Mario; Sipilae, Mikko; Stozhkov, Yuri; Stratmann, Frank; Tome, Antonio; Troestl, Jasmin; Tsagkogeorgas, Georgios; Vaattovaara, Petri; Viisanen, Yrjo; Virtanen, Annele; Vrtala, Aron; Wagner, Paul E.; Weingartner, Ernest; Wex, Heike; Williamson, Christina; Wimmer, Daniela; Ye, Penglin; Yli-Juuti, Taina; Carslaw, Kenneth S.; Kulmala, Markku; Curtius, Joachim; Baltensperger, Urs; Worsnop, Douglas R.; Vehkamaeki, Hanna; Kirkby, JasperNature (London, United Kingdom) (2013), 502 (7471), 359-363CODEN: NATUAS; ISSN:0028-0836. (Nature Publishing Group)Aerosol particle nucleation from trace atm. vapor is thought to provide up to half the global cloud condensation nuclei. Aerosols can cause net climate cooling by scattering sunlight and leading to smaller, but more numerous cloud droplets, making clouds brighter and extending their lifetimes. Atm. aerosols from human activities are thought to have compensated for a large fraction of warming caused by greenhouse gases; however, despite its importance for climate, atm. nucleation is poorly understood. It was recently shown that H2SO4 and NH3 cannot explain particle formation rates obsd. in the lower atm. It is thought amines may enhance nucleation, but until now there has been no direct evidence for amine ternary nucleation under atm. conditions. This work used the CLOUD (cosmics leaving outdoor droplets) chamber at CERN to det. that dimethylamine >3 parts per trillion by vol. can enhance particle formation rates >1000-fold vs. NH3, sufficient to account for atm. obsd. particle formation rates. Mol. anal. of clusters showed that faster nucleation is explained by a base-stabilization mechanism involving acid-amine pairs, which strongly decrease evapn. The ion-induced contribution is generally small, reflecting the high stability of H2SO4-dimethylamine clusters and indicating that galactic cosmic rays exert only a small effect on their formation, except at low overall formation rates. Exptl. measurements were well reproduced by a dynamic model based on quantum chem. calcns. of mol. cluster binding energies without any fitted parameters. Results showed that in regions of the atm. near amine sources, amines and SO2 should be considered when assessing the effect of anthropogenic activity on particle formation.
- 11Kurten, A.; Li, C.; Bianchi, F.; Curtius, J.; Dias, A.; Donahue, N. M.; Duplissy, J.; Flagan, R. C.; Hakala, J.; Jokinen, T. New particle formation in the sulfuric acid-dimethylamine-water system: Reevaluation of CLOUD chamber measurements and comparison to an aerosol nucleation and growth model. Atmospheric Chemistry and Physics 2018, 18, 845– 863, DOI: 10.5194/acp-18-845-2018There is no corresponding record for this reference.
- 12Rong, H.; Liu, J.; Zhang, Y.; Du, L.; Zhang, X.; Li, Z. Nucleation mechanisms of iodic acid in clean and polluted coastal regions. Chemosphere 2020, 253, 126743, DOI: 10.1016/j.chemosphere.2020.12674312https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXnsFait7s%253D&md5=661c4cafce75612ebd720a3f33d3416bNucleation mechanisms of iodic acid in clean and polluted coastal regionsRong, Hui; Liu, Jiarong; Zhang, Yujia; Du, Lin; Zhang, Xiuhui; Li, ZeshengChemosphere (2020), 253 (), 126743CODEN: CMSHAF; ISSN:0045-6535. (Elsevier Ltd.)In coastal regions, intense bursts of particles are frequently obsd. with high concns. of iodine species, esp. iodic acid (IA). However, the nucleation mechanisms of IA, esp. in polluted environments with high concns. of sulfuric acid (SA) and ammonia (A), remain to be fully established. By quantum chem. calcns. and atm. cluster dynamics code (ACDC) simulations, the self-nucleation of IA in clean coastal regions and that influenced by SA and A in polluted coastal regions are investigated. The results indicate that IA can form stable clusters stabilized by halogen bonds and hydrogen bonds through sequential addn. of IA, and the self-nucleation of IA can instantly produce large amts. of stable clusters when the concn. of IA is high during low tide, which is consistent with the observation that intense particle bursts were linked to high concns. of IA in clean coastal regions. Besides, SA and A can stabilize IA clusters by the formation of more halogen bonds and hydrogen bonds as well as proton transfers, and the binary nucleation of IA-SA/A rather than the self-nucleation of IA appears to be the dominant pathways in polluted coastal regions, esp. in winter. These new insights are helpful to understand the mechanisms of new particle formation induced by IA in clean and polluted coastal regions.
- 13He, X.-C.; Tham, Y. J.; Dada, L.; Wang, M.; Finkenzeller, H.; Stolzenburg, D.; Iyer, S.; Simon, M.; Kürten, A.; Shen, J. Role of iodine oxoacids in atmospheric aerosol nucleation. Science 2021, 371, 589– 595, DOI: 10.1126/science.abe029813https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXjsFCrtr8%253D&md5=f7b67dfba68a1d92579197f81529327bRole of iodine oxoacids in atmospheric aerosol nucleationHe, Xu-Cheng; Tham, Yee Jun; Dada, Lubna; Wang, Mingyi; Finkenzeller, Henning; Stolzenburg, Dominik; Iyer, Siddharth; Simon, Mario; Kurten, Andreas; Shen, Jiali; Rorup, Birte; Rissanen, Matti; Schobesberger, Siegfried; Baalbaki, Rima; Wang, Dongyu S.; Koenig, Theodore K.; Jokinen, Tuija; Sarnela, Nina; Beck, Lisa J.; Almeida, Joao; Amanatidis, Stavros; Amorim, Antonio; Ataei, Farnoush; Baccarini, Andrea; Bertozzi, Barbara; Bianchi, Federico; Brilke, Sophia; Caudillo, Lucia; Chen, Dexian; Chiu, Randall; Chu, Biwu; Dias, Antonio; Ding, Aijun; Dommen, Josef; Duplissy, Jonathan; El Haddad, Imad; Gonzalez Carracedo, Loic; Granzin, Manuel; Hansel, Armin; Heinritzi, Martin; Hofbauer, Victoria; Junninen, Heikki; Kangasluoma, Juha; Kemppainen, Deniz; Kim, Changhyuk; Kong, Weimeng; Krechmer, Jordan E.; Kvashin, Aleksander; Laitinen, Totti; Lamkaddam, Houssni; Lee, Chuan Ping; Lehtipalo, Katrianne; Leiminger, Markus; Li, Zijun; Makhmutov, Vladimir; Manninen, Hanna E.; Marie, Guillaume; Marten, Ruby; Mathot, Serge; Mauldin, Roy L.; Mentler, Bernhard; Mohler, Ottmar; Muller, Tatjana; Nie, Wei; Onnela, Antti; Petaja, Tuukka; Pfeifer, Joschka; Philippov, Maxim; Ranjithkumar, Ananth; Saiz-Lopez, Alfonso; Salma, Imre; Scholz, Wiebke; Schuchmann, Simone; Schulze, Benjamin; Steiner, Gerhard; Stozhkov, Yuri; Tauber, Christian; Tome, Antonio; Thakur, Roseline C.; Vaisanen, Olli; Vazquez-Pufleau, Miguel; Wagner, Andrea C.; Wang, Yonghong; Weber, Stefan K.; Winkler, Paul M.; Wu, Yusheng; Xiao, Mao; Yan, Chao; Ye, Qing; Ylisirnio, Arttu; Zauner-Wieczorek, Marcel; Zha, Qiaozhi; Zhou, Putian; Flagan, Richard C.; Curtius, Joachim; Baltensperger, Urs; Kulmala, Markku; Kerminen, Veli-Matti; Kurten, Theo; Donahue, Neil M.; Volkamer, Rainer; Kirkby, Jasper; Worsnop, Douglas R.; Sipila, MikkoScience (Washington, DC, United States) (2021), 371 (6529), 589-595CODEN: SCIEAS; ISSN:1095-9203. (American Association for the Advancement of Science)Iodic acid (HIO3) is known to form aerosol particles in coastal marine regions, but predicted nucleation and growth rates are lacking. Using the CERN CLOUD (Cosmics Leaving Outdoor Droplets) chamber, we find that the nucleation rates of HIO3 particles are rapid, even exceeding sulfuric acid-ammonia rates under similar conditions. We also find that ion-induced nucleation involves IO3- and the sequential addn. of HIO3 and that it proceeds at the kinetic limit below +10°C. In contrast, neutral nucleation involves the repeated sequential addn. of iodous acid (HIO2) followed by HIO3, showing that HIO2 plays a key stabilizing role. Freshly formed particles are composed almost entirely of HIO3, which drives rapid particle growth at the kinetic limit. Our measurements indicate that iodine oxoacid particle formation can compete with sulfuric acid in pristine regions of the atm.
- 14Baccarini, A.; Karlsson, L.; Dommen, J.; Duplessis, P.; Vullers, J.; Brooks, I. M.; Saiz-Lopez, A.; Salter, M.; Tjernstrom, M.; Baltensperger, U. Frequent new particle formation over the high Arctic pack ice by enhanced iodine emissions. Nat. Commun. 2020, 11, 4924, DOI: 10.1038/s41467-020-19533-y14https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXhvFyqt7jM&md5=8ce5a303b90658edab0b26febba072cbFrequent new particle formation over the high Arctic pack ice by enhanced iodine emissionsBaccarini, Andrea; Karlsson, Linn; Dommen, Josef; Duplessis, Patrick; Vullers, Jutta; Brooks, Ian M.; Saiz-Lopez, Alfonso; Salter, Matthew; Tjernstrom, Michael; Baltensperger, Urs; Zieger, Paul; Schmale, JuliaNature Communications (2020), 11 (1), 4924CODEN: NCAOBW; ISSN:2041-1723. (Nature Research)In the central Arctic Ocean the formation of clouds and their properties are sensitive to the availability of cloud condensation nuclei (CCN). The vapors responsible for new particle formation (NPF), potentially leading to CCN, have remained unidentified since the first aerosol measurements in 1991. Here, we report that all the obsd. NPF events from the Arctic Ocean 2018 expedition are driven by iodic acid with little contribution from sulfuric acid. Iodic acid largely explains the growth of ultrafine particles (UFP) in most events. The iodic acid concn. increases significantly from summer towards autumn, possibly linked to the ocean freeze-up and a seasonal rise in ozone. This leads to a one order of magnitude higher UFP concn. in autumn. Measurements of cloud residuals suggest that particles smaller than 30 nm in diam. can activate as CCN. Therefore, iodine NPF has the potential to influence cloud properties over the Arctic Ocean.
- 15Jimenez, J. L.; Canagaratna, M. R.; Donahue, N. M.; Prevot, A. S. H.; Zhang, Q.; Kroll, J. H.; DeCarlo, P. F.; Allan, J. D.; Coe, H.; Ng, N. L. Evolution of Organic Aerosols in the Atmosphere. Science (New York, N.Y.) 2009, 326, 1525– 9, DOI: 10.1126/science.118035315https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhsFensbjE&md5=dd5505995c2591c0540180493b0020eeEvolution of Organic Aerosols in the AtmosphereJimenez, J. L.; Canagaratna, M. R.; Donahue, N. M.; Prevot, A. S. H.; Zhang, Q.; Kroll, J. H.; DeCarlo, P. F.; Allan, J. D.; Coe, H.; Ng, N. L.; Aiken, A. C.; Docherty, K. S.; Ulbrich, I. M.; Grieshop, A. P.; Robinson, A. L.; Duplissy, J.; Smith, J. D.; Wilson, K. R.; Lanz, V. A.; Hueglin, C.; Sun, Y. L.; Tian, J.; Laaksonen, A.; Raatikainen, T.; Rautiainen, J.; Vaattovaara, P.; Ehn, M.; Kulmala, M.; Tomlinson, J. M.; Collins, D. R.; Cubison, M. J.; Dunlea, J.; Huffman, J. A.; Onasch, T. B.; Alfarra, M. R.; Williams, P. I.; Bower, K.; Kondo, Y.; Schneider, J.; Drewnick, F.; Borrmann, S.; Weimer, S.; Demerjian, K.; Salcedo, D.; Cottrell, L.; Griffin, R.; Takami, A.; Miyoshi, T.; Hatakeyama, S.; Shimono, A.; Sun, J. Y.; Zhang, Y. M.; Dzepina, K.; Kimmel, J. R.; Sueper, D.; Jayne, J. T.; Herndon, S. C.; Trimborn, A. M.; Williams, L. R.; Wood, E. C.; Middlebrook, A. M.; Kolb, C. E.; Baltensperger, U.; Worsnop, D. R.Science (Washington, DC, United States) (2009), 326 (5959), 1525-1529CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)Org. aerosol (OA) particles affect climate forcing and human health, but their sources and evolution are poorly characterized. A unifying model framework describing the atm. evolution of OA which is constrained by high time resolved measurements of its compn., volatility, and oxidn. state is presented. OA and OA precursor gases evolve by becoming increasingly oxidized, less volatile, and more hygroscopic, leading to the formation of oxygenated org. aerosols (OOA), with concns. comparable to those of SO42- aerosols throughout the Northern Hemisphere. This model framework captures the dynamic aging behavior obsd. in the atm. and lab.; it serves as a basis to improve regional and global model parameterizations.
- 16Petaja, T.; Tabakova, K.; Manninen, A.; Ezhova, E.; O’Connor, E.; Moisseev, D.; Sinclair, V. A.; Backman, J.; Levula, J.; Luoma, K. Influence of biogenic emissions from boreal forests on aerosol-cloud interactions. Nature Geoscience 2022, 15, 42– 47, DOI: 10.1038/s41561-021-00876-0There is no corresponding record for this reference.
- 17Tröstl, J.; Chuang, W.; Gordon, H.; Heinritzi, M.; Yan, C.; Molteni, U.; Ahlm, L.; Frege, C.; Bianchi, F.; Wagner, R. The role of low-volatility organic compounds in initial particle growth in the atmosphere. Nature 2016, 533, 527– 531, DOI: 10.1038/nature1827117https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28Xoslyrur4%253D&md5=7e09ef77e9daeb1f54cb52383f4feb68The role of low-volatility organic compounds in initial particle growth in the atmosphereTrostl, Jasmin; Chuang, Wayne K.; Gordon, Hamish; Heinritzi, Martin; Yan, Chao; Molteni, Ugo; Ahlm, Lars; Frege, Carla; Bianchi, Federico; Wagner, Robert; Simon, Mario; Lehtipalo, Katrianne; Williamson, Christina; Craven, Jill S.; Duplissy, Jonathan; Adamov, Alexey; Almeida, Joao; Bernhammer, Anne-Kathrin; Breitenlechner, Martin; Brilke, Sophia; Dias, Antonio; Ehrhart, Sebastian; Flagan, Richard C.; Franchin, Alessandro; Fuchs, Claudia; Guida, Roberto; Gysel, Martin; Hansel, Armin; Hoyle, Christopher R.; Jokinen, Tuija; Junninen, Heikki; Kangasluoma, Juha; Keskinen, Helmi; Kim, Jaeseok; Krapf, Manuel; Kurten, Andreas; Laaksonen, Ari; Lawler, Michael; Leiminger, Markus; Mathot, Serge; Mohler, Ottmar; Nieminen, Tuomo; Onnela, Antti; Petaja, Tuukka; Piel, Felix M.; Miettinen, Pasi; Rissanen, Matti P.; Rondo, Linda; Sarnela, Nina; Schobesberger, Siegfried; Sengupta, Kamalika; Sipila, Mikko; Smith, James N.; Steiner, Gerhard; Tome, Antonio; Virtanen, Annele; Wagner, Andrea C.; Weingartner, Ernest; Wimmer, Daniela; Winkler, Paul M.; Ye, Penglin; Carslaw, Kenneth S.; Curtius, Joachim; Dommen, Josef; Kirkby, Jasper; Kulmala, Markku; Riipinen, Ilona; Worsnop, Douglas R.; Donahue, Neil M.; Baltensperger, UrsNature (London, United Kingdom) (2016), 533 (7604), 527-531CODEN: NATUAS; ISSN:0028-0836. (Nature Publishing Group)About half of present-day cloud condensation nuclei originate from atm. nucleation, frequently appearing as a burst of new particles near midday. Atm. observations show that the growth rate of new particles often accelerates when the diam. of the particles is between one and ten nanometers. In this crit. size range, new particles are most likely to be lost by coagulation with pre-existing particles, thereby failing to form new cloud condensation nuclei that are typically 50 to 100 nm across. Sulfuric acid vapor is often involved in nucleation but is too scarce to explain most subsequent growth, leaving org. vapors as the most plausible alternative, at least in the planetary boundary layer. Although recent studies predict that low-volatility org. vapors contribute during initial growth, direct evidence has been lacking. The accelerating growth may result from increased photolytic prodn. of condensable org. species in the afternoon, and the presence of a possible Kelvin (curvature) effect, which inhibits org. vapor condensation on the smallest particles (the nano-Kohler theory), has so far remained ambiguous. The authors presented expts. performed in a large chamber under atm. conditions that investigate the role of org. vapors in the initial growth of nucleated org. particles in the absence of inorg. acids and bases such as sulfuric acid or ammonia and amines, resp. Using data from the same set of expts., it has been shown that org. vapors alone can drive nucleation. They focused on the growth of nucleated particles and find that the org. vapors that drive initial growth have extremely low volatilities (satn. concn. less than 10-4.5 micrograms per cubic meter). As the particles increase in size and the Kelvin barrier falls, subsequent growth is primarily due to more abundant org. vapors of slightly higher volatility (satn. concns. of 10-4.5 to 10-0.5 micrograms per cubic meter). They present a particle growth model that quant. reproduces measurements. They implement a parameterization of the first steps of growth in a global aerosol model and find that concns. of atm. cloud concn. nuclei can change substantially in response, i.e., by up to 50 per cent in comparison with previously assumed growth rate parameterizations.
- 18Crounse, J.; Nielsen, L.; Jørgensen, S.; Kjaergaard, H.; Wennberg, P. Autoxidation of Organic Compounds in the Atmosphere. JOURNAL OF PHYSICAL. CHEMISTRY LETTERS 2013, 4, 3513– 3520, DOI: 10.1021/jz4019207There is no corresponding record for this reference.
- 19Bianchi, F.; Kurtén, T.; Riva, M.; Mohr, C.; Rissanen, M.; Roldin, P.; Berndt, T.; Crounse, J.; Wennberg, P.; Mentel, T.; Wildt, J.; Junninen, H.; Jokinen, T.; Kulmala, M.; Worsnop, D.; Thornton, J.; Donahue, N.; Kjaergaard, H.; Ehn, M. Highly Oxygenated Molecules (HOM) from Gas-Phase Autoxidation Involving Organic Peroxy Radicals: A Key Contributor to Atmospheric Aerosol. Chem. Rev. 2019, 119, 3472– 3509, DOI: 10.1021/acs.chemrev.8b0039519https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXjsFSrsb0%253D&md5=860f5dce6415fbbdee9f45b7afeff559Highly Oxygenated Molecules (HOM) from Gas-Phase Autoxidation Involving Organic Peroxy Radicals: A Key Contributor to Atmospheric AerosolBianchi, Federico; Kurten, Theo; Riva, Matthieu; Mohr, Claudia; Rissanen, Matti P.; Roldin, Pontus; Berndt, Torsten; Crounse, John D.; Wennberg, Paul O.; Mentel, Thomas F.; Wildt, Jurgen; Junninen, Heikki; Jokinen, Tuija; Kulmala, Markku; Worsnop, Douglas R.; Thornton, Joel A.; Donahue, Neil; Kjaergaard, Henrik G.; Ehn, MikaelChemical Reviews (Washington, DC, United States) (2019), 119 (6), 3472-3509CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review which defines highly oxygenated org. mols. (HOM) formed in the atm. via auto-oxidn. involving peroxy radicals arising from volatile org. compds. describing currently available techniques for their identification/quantification, followed by a summary of the current knowledge on their formation mechanisms and physicochem. properties, is given. Major aims are to provide a common frame for the currently quite fragmented literature on HOM studies and highlighting existing gaps, and suggesting directions for future HOM research. Topics discussed include: introduction; HOM background (defining key concepts, HOM in relation to other classification schemes, historical naming conventions); HOM detection (gas and particle phases, uncertainties and anal. challenges of HOM detection); HOM formation mechanisms (auto-oxidn. involving peroxy radical as HOM source, bimol. RO2 reactions, factors affecting HOM formation); HOM properties and atm. fate (physicochem. properties, removal mechanisms); HOM atm. observations and impact (ambient HOM observation, atm. impact); and summary and perspectives.
- 20Roldin, P.; Ehn, M.; Kurten, T.; Olenius, T.; Rissanen, M. P.; Sarnela, N.; Elm, J.; Rantala, P.; Hao, L.; Hyttinen, N. The role of highly oxygenated organic molecules in the Boreal aerosol-cloud-climate system. Nat. Commun. 2019, 10, 1– 15, DOI: 10.1038/s41467-019-12338-8There is no corresponding record for this reference.
- 21Xavier, C.; de jonge, R. W.; Jokinen, T.; Beck, L.; Sipilä, M.; Olenius, T.; Roldin, P. Role of Iodine-Assisted Aerosol Particle Formation in Antarctica. Environ. Sci. Technol. 2024, 58, 7314– 7324, DOI: 10.1021/acs.est.3c09103There is no corresponding record for this reference.
- 22Brean, J.; Dall’Osto, M.; Simo, R.; Shi, Z.; Beddows, D. C. S.; Harrison, R. M. Open ocean and coastal new particle formation from sulfuric acid and amines around the Antarctic Peninsula. Nature Geoscience 2021, 14, 383– 388, DOI: 10.1038/s41561-021-00751-yThere is no corresponding record for this reference.
- 23Jokinen, T.; Sipila, M.; Kontkanen, J.; Vakkari, V.; Tisler, P.; Duplissy, E.-M.; Junninen, H.; Kangasluoma, J.; Manninen, H. E.; Petaja, T. Ion-induced sulfuric acid-ammonia nucleation drives particle formation in coastal Antarctica. Science Advances 2018, 4, eaat9744, DOI: 10.1126/sciadv.aat9744There is no corresponding record for this reference.
- 24Lee, H.; Lee, K.; Lunder, C.; Krejci, R.; Aas, W.; Jiyeon, P.; Park, K.-t.; Lee, B.; Yoon, Y.-J.; Park, K. Atmospheric new particle formation characteristics in the Arctic as measured at Mount Zeppelin, Svalbard, from 2016 to 2018. Atmospheric Chem. Phys. 2020, 20 (21), 13425– 13441There is no corresponding record for this reference.
- 25Kecorius, S.; Vogl, T.; Paasonen, P.; Lampilahti, J.; Rothenberg, D.; Wex, H.; Zeppenfeld, S.; van Pinxteren, M.; Hartmann, M.; Henning, S. New particle formation and its effect on cloud condensation nuclei abundance in the summer Arctic: a case study in the Fram Strait and Barents Sea. Atmospheric Chemistry and Physics 2019, 19, 14339– 14364, DOI: 10.5194/acp-19-14339-2019There is no corresponding record for this reference.
- 26DallÓsto, M.; Geels, C.; Beddows, D. C. S.; Boertmann, D.; Lange, R.; Nøjgaard, J. K.; Harrison, R. M.; Simo, R.; Skov, H.; Massling, A. Regions of open water and melting sea ice drive new particle formation in North East Greenland. Sci. Rep. 2018, 8, 6109, DOI: 10.1038/s41598-018-24426-8There is no corresponding record for this reference.
- 27Mäkelä, J.; Yli-Koivisto, S.; Hiltunen, V.; Seidl, W.; Swietlicki, E.; Teinilä, K.; Sillanpää, M.; Koponen, I.; Paatero, J.; Rosman, K.; Hämeri, K. Chemical composition of aerosol during particle formation events in boreal forest. Tellus B 2001, 53, 380– 393, DOI: 10.1034/j.1600-0889.2001.530405.x27https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXns1Gktrs%253D&md5=ce371260e6101c15c4bbaac0a5473bb1Chemical composition of aerosol during particle formation events in boreal forestMakela, J. M.; Yli-Koivisto, S.; Hiltunen, V.; Seidl, W.; Swietlicki, E.; Teinila, K.; Sillanpaa, M.; Koponen, I. K.; Paatero, J.; Rosman, K.; Hameri, K.Tellus, Series B: Chemical and Physical Meteorology (2001), 53B (4), 380-393CODEN: TSBMD7; ISSN:0280-6509. (Munksgaard International Publishers Ltd.)Size-segregated chem. aerosol anal. of 5 integrated samples was performed for atm. aerosols during new particle formation events in spring 1999 during the BIOFOR 3 measurement campaign at a boreal forest site in southern Finland. Aerosol samples collected by a cascade low-pressure impactor were taken selectively to distinguish particle formation event aerosols from non-event aerosols. The division into event and non-event cases was done in-situ in the field, based on the online submicron no. size distribution. Results of chem. ionic compn. of particles showed only small differences between event and non-event sample sets. Event samples had lower concns. of total SO42- and NH4+ and light dicarboxylic acids, e.g., oxalate, malonate, and succinate. In event samples, nucleation mode particle MSA (methane sulfonic acid) was present in concns. exceeding those obsd. in non-event samples; however, at larger particle sizes, sample sets contained rather similar MSA concns. The most significant difference between event and non-event sets was obsd. for dimethylammonium, the ionic component of dimethylamine ((CH3)2NH), which seemed to be present in the particle phase during particle formation periods and/or during subsequent particle growth. The abs. event sample dimethylamine concns. were >30-fold greater than non-event concns. in the accumulation mode size range. The non-event back-up filter stage for sub-30 nm particles contained more dimethylamine than event samples. This fractionation is probably a condensation artifact from impactor sampling. A simple mass balance est. was performed to evaluate the quality and consistency of results for overall mass concn.
- 28Hemmilä, M.; Hellén, H.; Virkkula, A.; Makkonen, U.; Praplan, A.; Kontkanen, J.; Ahonen, L.; Kulmala, M.; Hakola, H. Amines in boreal forest air at SMEAR II station in Finland. Atmospheric Chemistry and Physics 2018, 18, 6367– 6380, DOI: 10.5194/acp-18-6367-2018There is no corresponding record for this reference.
- 29Lawler, M.; Rissanen, M.; Ehn, M.; Mauldin, R.; Sarnela, N.; Sipilä, M.; Smith, J. Evidence for Diverse Biogeochemical Drivers of Boreal Forest New Particle Formation. Geophys. Res. Lett. 2018, 45, 2038– 2046, DOI: 10.1002/2017GL076394There is no corresponding record for this reference.
- 30Sogacheva, L.; Saukkonen, L.; Nilsson, E.; Dal Maso, M.; Schultz, D.; de Leeuw, G.; Kulmala, M. New aerosol particle formation in different synoptic situations at Hyytiälä, Southern Finland. Tellus B 2022, 60, 485– 494, DOI: 10.1111/j.1600-0889.2008.00364.xThere is no corresponding record for this reference.
- 31Nieminen, T.; Yli-Juuti, T.; Manninen, H. E.; Petäjä, T.; Kerminen, V.-M.; Kulmala, M. Technical note: New particle formation event forecasts during PEGASOS-Zeppelin Northern mission 2013 in Hyytiälä, Finland. Atmospheric Chemistry and Physics 2015, 15, 12385– 12396, DOI: 10.5194/acp-15-12385-2015There is no corresponding record for this reference.
- 32Öström, E.; Putian, Z.; Schurgers, G.; Mishurov, M.; Kivekäs, N.; Lihavainen, H.; Ehn, M.; Rissanen, M. P.; Kurtén, T.; Boy, M.; Swietlicki, E.; Roldin, P. Modeling the role of highly oxidized multifunctional organic molecules for the growth of new particles over the boreal forest region. Atmospheric Chemistry and Physics 2017, 17, 8887– 8901, DOI: 10.5194/acp-17-8887-2017There is no corresponding record for this reference.
- 33Jenkin, M. E.; Saunders, S. M.; Pilling, M. J. The tropospheric degradation of volatile organic compounds: a protocol for mechanism development. Atmos. Environ. 1997, 31, 81– 104, DOI: 10.1016/S1352-2310(96)00105-733https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XntVOnt78%253D&md5=44dc875e8e1a9456ed2b570393d450c6The tropospheric degradation of volatile organic compounds: a protocol for mechanism developmentJenkin, Michael E.; Saunders, Sandra M.; Pilling, Michael J.Atmospheric Environment (1996), 31 (1), 81-104CODEN: AENVEQ; ISSN:1352-2310. (Elsevier)Kinetic and mechanistic data relevant to the tropospheric oxidn. of volatile org. compds. (VOCs) were used to define a series of rules for the construction of detailed degrdn. schemes for use in numerical models. These rules are intended to apply to the treatment of a wide range of non-arom. hydrocarbons and oxygenated and chlorinated VOCs, and are currently used to provide an up-to-date mechanism describing the degrdn. of a range of VOCs, and the formation of secondary oxidants, for use in a model of the boundary layer over Europe. The schemes constructed using this protocol are applicable, however, to a wide range of ambient conditions, and may be employed in models of urban, rural, or remote tropospheric environments, or for the simulation of secondary pollutant formation for a range of NOx or VOC emission scenarios. These schemes are believed to be particularly appropriate for comparative assessments of the formation of oxidants, such as ozone, from the degrdn. of org. compds. The protocol is divided into a series of subsections dealing with initiation reactions, the reactions of the radical intermediates and the further degrdn. of first and subsequent generation products. The present work draws heavily on previous reviews and evaluations of data relevant to tropospheric chem. Where necessary, however, existing recommendations are adapted, or new rules are defined, to reflect recent improvements in the database, particularly with regard to the treatment of peroxy radical (RO2) reactions for which there have been major advances, even since comparatively recent reviews. The present protocol aims to take into consideration work available in the open literature up to the end of 1994, and some further studies known by the authors, which were under review at that time. A major disadvantage of explicit chem. mechanisms is the very large no. of reactions potentially generated, if a series of rules is rigorously applied. The protocol aims to limit the no. of reactions in a degrdn. scheme by applying a degree of strategic simplication, while maintaining the essential features of the chem. These simplication measures are described, and their influence is demonstrated and discussed. The resultant mechanisms are believed to provide a suitable starting point for the generation of reduced chem. mechanisms.
- 34Jenkin, M. E.; Saunders, S. M.; Wagner, V.; Pilling, M. J. Protocol for the development of the Master Chemical Mechanism, MCM v3 (Part B): tropospheric degradation of aromatic volatile organic compounds. Atmospheric Chemistry and Physics 2003, 3, 181– 193, DOI: 10.5194/acp-3-181-200334https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXns1Sntrw%253D&md5=cbb782c7d9ee66115b9c13f14758d095Protocol for the development of the master chemical mechanism, MCM v3 (part B): tropospheric degradation of aromatic volatile organic compoundsJenkin, M. E.; Saunders, S. M.; Wagner, V.; Pilling, M. J.Atmospheric Chemistry and Physics (2003), 3 (1), 181-193CODEN: ACPTCE; ISSN:1680-7324. (European Geophysical Society)Kinetic and mechanistic data relevant to the tropospheric degrdn. of arom. volatile org. compds. (VOC) were used to define a mechanism development protocol, which was used to construct degrdn. schemes for 18 arom. VOC as part of version 3 of the Master Chem. Mechanism (MCM v3). This is complementary to the treatment of 107 nonarom. VOC, presented in a companion paper. The protocol is divided into subsections describing initiation reactions, the degrdn. chem. to 1st generation products via a no. of competitive routes, and the further degrdn. of 1st and subsequent generation products. Emphasis is placed on describing where the treatment differs from that applied to the nonarom. VOC. The protocol is based on work available in the open literature up to the beginning of 2001, and some other studies known by the authors which were under review at the time. Photochem. Ozone Creation Potentials (POCP) were calcd. for the 18 arom. VOC in MCM v3 for idealized conditions appropriate to north-west Europe, using a photochem. trajectory model. The POCP values provide a measure of the relative ozone forming abilities of the VOC. These show distinct differences from POCP values calcd. previously for the aroms., using earlier versions of the MCM, and reasons for these differences are discussed.
- 35Braeuer, P.; Tilgner, A.; Wolke, R.; Herrmann, H. Mechanism development and modelling of tropospheric multiphase halogen chemistry: The CAPRAM Halogen Module 2.0 (HM2). JOURNAL OF ATMOSPHERIC CHEMISTRY 2013, 70, 19– 52, DOI: 10.1007/s10874-013-9249-6There is no corresponding record for this reference.
- 36Wu, R.; Wang, S.; Wang, L. A New Mechanism for The Atmospheric Oxidation of Dimethyl Sulfide. The Importance of Intramolecular Hydrogen Shift in CH3SCH2OO Radical. journal of physical chemistry. A 2015, 119, 112, DOI: 10.1021/jp511616jThere is no corresponding record for this reference.
- 37Berndt, T.; Scholz, W.; Mentler, B.; Fischer, L.; Hoffmann, E.; Tilgner, A.; Hyttinen, N.; Prisle, N. L.; Hansel, A.; Herrmann, H. Fast Peroxy Radical Isomerization and OH Recycling in the Reaction of OH Radicals with Dimethyl Sulfide. J. Phys. Chem. Lett. 2019, 10, 6478, DOI: 10.1021/acs.jpclett.9b0256737https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXhvFWltrjP&md5=0d91ffa9cadc1365ffbfb0a48fe0373bFast Peroxy Radical Isomerization and OH Recycling in the Reaction of OH Radicals with Dimethyl SulfideBerndt, T.; Scholz, W.; Mentler, B.; Fischer, L.; Hoffmann, E. H.; Tilgner, A.; Hyttinen, N.; Prisle, N. L.; Hansel, A.; Herrmann, H.Journal of Physical Chemistry Letters (2019), 10 (21), 6478-6483CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)Di-Me sulfide (DMS), produced by marine organisms, represents the most abundant, biogenic sulfur emission into the Earth's atm. The gas-phase degrdn. of DMS is mainly initiated by the reaction with the OH radical forming first CH3SCH2O2 radicals from the dominant H-abstraction channel. It is exptl. shown that these peroxy radicals undergo a two-step isomerization process finally forming a product consistent with the formula HOOCH2SCHO. The isomerization process is accompanied by OH recycling. The rate-limiting first isomerization step, CH3SCH2O2 → CH2SCH2OOH, followed by O2 addn., proceeds with k = (0.23 ± 0.12) s-1 at 295 ± 2 K. Competing bimol. CH3SCH2O2 reactions with NO, HO2, or RO2 radicals are less important for trace-gas conditions over the oceans. Results of atm. chem. simulations demonstrate the predominance (≥95%) of CH3SCH2O2 isomerization. The rapid peroxy radical isomerization, not yet considered in models, substantially changes the understanding of DMS's degrdn. processes in the atm.
- 38Veres, P. R.; Neuman, J. A.; Bertram, T. H.; Assaf, E.; Wolfe, G. M.; Williamson, C. J.; Weinzierl, B.; Tilmes, S.; Thompson, C. R.; Thames, A. B.; Schroder, J. C.; Saiz-Lopez, A.; Rollins, A. W.; Roberts, J. M.; Price, D.; Peischl, J.; Nault, B. A.; Møller, K. H.; Miller, D. O.; Meinardi, S.; Li, Q.; Lamarque, J.-F.; Kupc, A.; Kjaergaard, H. G.; Kinnison, D.; Jimenez, J. L.; Jernigan, C. M.; Hornbrook, R. S.; Hills, A.; Dollner, M.; Day, D. A.; Cuevas, C. A.; Campuzano-Jost, P.; Burkholder, J.; Bui, T. P.; Brune, W. H.; Brown, S. S.; Brock, C. A.; Bourgeois, I.; Blake, D. R.; Apel, E. C.; Ryerson, T. B. Global airborne sampling reveals a previously unobserved dimethyl sulfide oxidation mechanism in the marine atmosphere. Proc. Natl. Acad. Sci. U. S. A. 2020, 117, 4505– 4510, DOI: 10.1073/pnas.191934411738https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXksVSqsr0%253D&md5=0d957baa667a2543cd9a4f3c467790a5Global airborne sampling reveals a previously unobserved dimethyl sulfide oxidation mechanism in the marine atmosphereVeres, Patrick R.; Neuman, J. Andrew; Bertram, Timothy H.; Assaf, Emmanuel; Wolfe, Glenn M.; Williamson, Christina J.; Weinzierl, Bernadett; Tilmes, Simone; Thompson, Chelsea R.; Thames, Alexander B.; Schroder, Jason C.; Saiz-Lopez, Alfonso; Rollins, Andrew W.; Roberts, James M.; Price, Derek; Peischl, Jeff; Nault, Benjamin A.; Moeller, Kristian H.; Miller, David O.; Meinardi, Simone; Li, Qinyi; Lamarque, Jean-Francois; Kupc, Agnieszka; Kjaergaard, Henrik G.; Kinnison, Douglas; Jimenez, Jose L.; Jernigan, Christopher M.; Hornbrook, Rebecca S.; Hills, Alan; Dollner, Maximilian; Day, Douglas A.; Cuevas, Carlos A.; Campuzano-Jost, Pedro; Burkholder, James; Bui, T. Paul; Brune, William H.; Brown, Steven S.; Brock, Charles A.; Bourgeois, Ilann; Blake, Donald R.; Apel, Eric C.; Ryerson, Thomas B.Proceedings of the National Academy of Sciences of the United States of America (2020), 117 (9), 4505-4510CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)Di-Me sulfide (DMS), emitted from the oceans, is the most abundant biol. source of sulfur to the marine atm. Atm. DMS is oxidized to condensable products that form secondary aerosols that affect Earth's radiative balance by scattering solar radiation and serving as cloud condensation nuclei. We report the atm. discovery of a previously unquantified DMS oxidn. product, hydroperoxymethyl thioformate (HPMTF, HOOCH2SCHO), identified through global-scale airborne observations that demonstrate it to be a major reservoir of marine sulfur. Observationally constrained model results show that more than 30% of oceanic DMS emitted to the atm. forms HPMTF. Coincident particle measurements suggest a strong link between HPMTF concn. and new particle formation and growth. Analyses of these observations show that HPMTF chem. must be included in atm. models to improve representation of key linkages between the biogeochem. of the ocean, marine aerosol formation and growth, and their combined effects on climate.
- 39Hari, P.; Kulmala, M. Station for Measuring Ecosystem-Atmosphere Relations (SMEAR II). Boreal Environment Research 2005, 10, 315– 32239https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXht1GhurvP&md5=fbd8465f76a998b3a0958c12089db374Station for Measuring Ecosystem-Atmosphere Relations (SMEAR II)Hari, Pertti; Kulmala, MarkkuBoreal Environment Research (2005), 10 (5), 315-322CODEN: BEREF7; ISSN:1239-6095. (Finnish Environment Institute)Here we present the ongoing SMEAR (Station for Measuring Forest Ecosystem-Atm. Relations) research program and also related future views. The main idea of SMEAR-type infrastructures is continuous, comprehensive measurements of fluxes, storages and concns. in the land ecosystem-atm. continuum. The major coupling mechanisms between atm. and land surface are the fluxes of energy, momentum, water, carbon dioxide, atm. trace gases and atm. aerosols. Understanding of couplings and feedbacks is the basis for the prediction of changes in the system formed by atm., vegetation and soil. A better quantification of the agents that cause climate change, as well as the emissions and removals of species, will provide more accurate projections of future atm. compn. and hence climate.
- 40Neefjes, I.; Laapas, M.; Liu, Y.; Médus, E.; Miettunen, E.; Ahonen, L.; Quéléver, L.; Aalto, J.; Bäck, J.; Kerminen, V.-M.; Lamplahti, J.; Luoma, K.; Maki, M.; Mammarella, I.; Petäjä, T.; Räty, M.; Sarnela, N.; Ylivinkka, I.; Hakala, S.; Lintunen, A. 25 years of atmospheric and ecosystem measurements in a boreal forestSeasonal variation and responses to warm and dry years. Boreal Environment Research 2022, 27, 1– 31There is no corresponding record for this reference.
- 41Stein, A. F.; Draxler, R. R.; Rolph, G. D.; Stunder, B. J. B.; Cohen, M. D.; Ngan, F. NOAAs HYSPLIT Atmospheric Transport and Dispersion Modeling System. Bulletin of the American Meteorological Society 2015, 96, 2059– 2077, DOI: 10.1175/BAMS-D-14-00110.1There is no corresponding record for this reference.
- 42Rolph, G.; Stein, A.; Stunder, B. Real-time Environmental Applications and Display sYstem: READY. Environmental Modelling & Software 2017, 95, 210– 228, DOI: 10.1016/j.envsoft.2017.06.025There is no corresponding record for this reference.
- 43Lennartz, S. T.; Marandino, C. A.; von Hobe, M.; Cortes, P.; Quack, B.; Simo, R.; Booge, D.; Pozzer, A.; Steinhoff, T.; Arevalo-Martinez, D. L. Direct oceanic emissions unlikely to account for the missing source of atmospheric carbonyl sulfide. Atmospheric Chemistry and Physics 2017, 17, 385– 402, DOI: 10.5194/acp-17-385-2017There is no corresponding record for this reference.
- 44Ziska, F.; Quack, B.; Abrahamsson, K.; Archer, S. D.; Atlas, E.; Bell, T.; Butler, J. H.; Carpenter, L. J.; Jones, C. E.; Harris, N. R. P. Global sea-to-air flux climatology for bromoform, dibromomethane and methyl iodide. Atmospheric Chemistry and Physics Discussions 2013, 13, 8915– 8934, DOI: 10.5194/acp-13-8915-201344https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhslehtrbL&md5=d13f982b82d94c732ab4dc017908098cGlobal sea-to-air flux climatology for bromoform, dibromomethane and methyl iodideZiska, F.; Quack, B.; Abrahamsson, K.; Archer, S. D.; Atlas, E.; Bell, T.; Butler, J. H.; Carpenter, L. J.; Jones, C. E.; Harris, N. R. P.; Hepach, H.; Heumann, K. G.; Hughes, C.; Kuss, J.; Krueger, K.; Liss, P.; Moore, R. M.; Orlikowska, A.; Raimund, S.; Reeves, C. E.; Reifenhaeuser, W.; Robinson, A. D.; Schall, C.; Tanhua, T.; Tegtmeier, S.; Turner, S.; Wang, L.; Wallace, D.; Williams, J.; Yamamoto, H.; Yvon-Lewis, S.; Yokouchi, Y.Atmospheric Chemistry and Physics (2013), 13 (17), 8915-8934, 20 pp.CODEN: ACPTCE; ISSN:1680-7324. (Copernicus Publications)Volatile halogenated org. compds. contg. bromine and iodine, which are naturally produced in the ocean, are involved in ozone depletion in both the troposphere and stratosphere. Three prominent compds. transporting large amts. of marine halogens into the atm. are bromoform (CHBr3), dibromomethane (CH2Br2) and Me iodide (CH3I). The input of marine halogens to the stratosphere has been estd. from observations and modeling studies using low-resoln. oceanic emission scenarios derived from top-down approaches. In order to improve emission inventory ests., we calc. data-based high resoln. global sea-to-air flux ests. of these compds. from surface observations within the HalOcAt (Halocarbons in the Ocean and Atm.) database. Global maps of marine and atm. surface concns. are derived from the data which are divided into coastal, shelf and open ocean regions. Considering phys. and biogeochem. characteristics of ocean and atm., the open ocean water and atm. data are classified into 21 regions. The available data are interpolated onto a 1° × 1° grid while missing grid values are interpolated with latitudinal and longitudinal dependent regression techniques reflecting the compds.' distributions. With the generated surface concn. climatologies for the ocean and atm., global sea-to-air concn. gradients and sea-to-air fluxes are calcd. Based on these calcns. we est. a total global flux of 1.5/2.5 Gmol Br yr-1 for CHBr3, 0.78/0.98 Gmol Br yr-1 for CH2Br2 and 1.24/1.45 Gmol Br yr-1 for CH3I (robust fit/ordinary least squares regression techniques). Contrary to recent studies, neg. fluxes occur in each sea-to-air flux climatol., mainly in the Arctic and Antarctic regions. "Hot spots" for global polybromomethane emissions are located in the equatorial region, whereas Me iodide emissions are enhanced in the subtropical gyre regions. Inter-annual and seasonal variation is contained within our flux calcns. for all three compds. Compared to earlier studies, our global fluxes are at the lower end of ests., esp. for bromoform. An under-representation of coastal emissions and of extreme events in our est. might explain the mismatch between our bottom-up emission est. and top-down approaches.
- 45Nightingale, P. D.; Malin, G.; Law, C. S.; Watson, A. J.; Liss, P. S.; Liddicoat, M. I.; Boutin, J.; Upstill-Goddard, R. C. In situ evaluation of air-sea gas exchange parameterizations using novel conservative and volatile tracers. Global Biogeochemical Cycles 2000, 14, 373– 387, DOI: 10.1029/1999GB90009145https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXhvVGms7s%253D&md5=d317f0ad2e899416471475672bd335f3In situ evaluation of air-sea gas exchange parameterizations using novel conservative and volatile tracersNightingale, Philip D.; Malin, Gill; Law, Cliff S.; Watson, Andrew J.; Liss, Peter S.; Liddicoat, Malcolm I.; Boutin, Jacqueline; Upstill-Goddard, Robert C.Global Biogeochemical Cycles (2000), 14 (1), 373-387CODEN: GBCYEP; ISSN:0886-6236. (American Geophysical Union)Measurements of air-sea gas exchange rates are reported from 2 deliberate tracer expts. in the southern North Sea during Feb. 1992 and 1993. A conservative tracer, spores of Bacillus globigii var. Niger, was used for the 1st time in an in situ air-sea gas exchange expt. This nonvolatile tracer is used to correct for dispersive diln. of the volatile tracers and allows 3 estns. of the transfer velocity for the same time period. The 1st estn. of the power dependence of gas transfer on mol. diffusivity in the marine environment is reported. This allows the impact of bubbles on ests. of the transfer velocity derived from changes in the He/SF6 ratio to be assessed. Data from earlier dual tracer expts. are reinterpreted, and findings suggest that results from all dual tracer expts. are mutually consistent. The complete data set is used to test published parameterizations of gas transfer with wind speed. A gas exchange relation that shows a dependence on wind speed intermediate between those of Liss and Merlivat (1986) and Wanninkhof (1992) is optimal. The dual tracer data are shown to be reasonably consistent with global ests. of gas exchange based on the uptake of natural and bomb-derived radiocarbon. The degree of scatter in the data when plotted against wind speed suggests that parameters not scaling with wind speed are also influencing gas exchange rates.
- 46Lana, A.; Bell, T. G.; Simó, R.; Vallina, S. M.; Ballabrera-Poy, J.; Kettle, A. J.; Dachs, J.; Bopp, L.; Saltzman, E. S.; Stefels, J. An updated climatology of surface dimethlysulfide concentrations and emission fluxes in the global ocean. Global Biogeochemical Cycles 2011, 25, 886, DOI: 10.1029/2010GB003850There is no corresponding record for this reference.
- 47Granier, C. S.; Darras, H.; Denier van der Gon, J.; Doubalova, N.; Elguindi, B.; Galle, M.; Gauss, M.; Guevara, J.; Jalkanen, J.; Kuenen, C.; Liousse, B.; Quack, D.; Simpson, K. The Copernicus Atmosphere Monitoring Service global and regional emissions. Sindelarova The Copernicus Atmosphere Monitoring Service global and regional emissions (April 2019 version) 2019, 16, DOI: 10.24380/d0bn-kx16There is no corresponding record for this reference.
- 48Sindelarova, K.; Granier, C.; Bouarar, I.; Guenther, A.; Tilmes, S.; Stavrakou, T.; Muller, J.-F.; Kuhn, U.; Stefani, P.; Knorr, W. Global dataset of biogenic VOC emissions calculated by the MEGAN model over the last 30 years. Atmospheric Chemistry and Physics 2014, 14, 9317, DOI: 10.5194/acp-14-9317-201448https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhs1yktbjO&md5=978d8465b40240c42347d621604c1516Global data set of biogenic VOC emissions calculated by the MEGAN model over the last 30 yearsSindelarova, K.; Granier, C.; Bouarar, I.; Guenther, A.; Tilmes, S.; Stavrakou, T.; Muller, J.-F.; Kuhn, U.; Stefani, P.; Knorr, W.Atmospheric Chemistry and Physics (2014), 14 (17), 9317-9341, 25 pp.CODEN: ACPTCE; ISSN:1680-7324. (Copernicus Publications)The Model of Emissions of Gases and Aerosols from Nature (MEGANv2.1) together with the Modern-Era Retrospective Anal. for Research and Applications (MERRA) meteorol. fields were used to create a global emission data set of biogenic volatile org. compds. (BVOC) available on a monthly basis for the time period of 1980-2010. This data set, developed under the Monitoring Atm. Compn. and Climate project (MACC), is called MEGAN-MACC. The model estd. mean annual total BVOC emission of 760 Tg (C) yr-1 consisting of isoprene (70%), monoterpenes (11%), methanol (6%), acetone (3%), sesquiterpenes (2.5%) and other BVOC species each contributing less than 2%. Several sensitivity model runs were performed to study the impact of different model input and model settings on isoprene ests. and resulted in differences of up to ±17% of the ref. isoprene total. A greater impact was obsd. for a sensitivity run applying parameterization of soil moisture deficit that led to a 50% redn. of isoprene emissions on a global scale, most significantly in specific regions of Africa, South America and Australia. MEGAN-MACC ests. are comparable to results of previous studies. More detailed comparison with other isoprene inventories indicated significant spatial and temporal differences between the data sets esp. for Australia, Southeast Asia and South America. MEGAN-MACC ests. of isoprene, α-pinene and group of monoterpenes showed a reasonable agreement with surface flux measurements at sites located in tropical forests in the Amazon and Malaysia. The model was able to capture the seasonal variation of isoprene emissions in the Amazon forest.
- 49Smith, B.; Wårlind, D.; Arneth, A.; Hickler, T.; Leadley, P.; Siltberg, J.; Zaehle, S. Implications of incorporating N cycling and N limitations on primary production in an individual-based dynamic vegetation model. Biogeosciences 2014, 11, 2027– 2054, DOI: 10.5194/bg-11-2027-2014There is no corresponding record for this reference.
- 50Arneth, A.; Niinemets, U.; Pressley, S.; Bäck, J.; Hari, P.; Karl, T.; Noe, S.; Prentice, I. C.; Serça, D.; Hickler, T.; Wolf, A.; Smith, B. Process-based estimates of terrestrial ecosystem isoprene emissions: incorporating the effects of a direct CO2-isoprene interaction. Atmospheric Chemistry and Physics 2007, 7, 31– 53, DOI: 10.5194/acp-7-31-2007There is no corresponding record for this reference.
- 51Paulot, F.; Jacob, D. J.; Johnson, M. T.; Bell, T. G.; Baker, A. R.; Keene, W. C.; Lima, I. D.; Doney, S. C.; Stock, C. A. Global oceanic emission of ammonia: Constraints from seawater and atmospheric observations. Global Biogeochemical Cycles 2015, 29, 1165– 1178, DOI: 10.1002/2015GB005106There is no corresponding record for this reference.
- 52Carpenter, L.; MacDonald, S.; Shaw, M.; Kumar, R.; Saunders, R.; Parthipan, R.; Wilson, J.; Plane, J. Atmospheric iodine levels influenced by sea surface emissions of inorganic iodine. Nature Geoscience 2013, 6, 108– 111, DOI: 10.1038/ngeo168752https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXntFyqsw%253D%253D&md5=fccd3f2ce8410029ba975238eb11da97Atmospheric iodine levels influenced by sea surface emissions of inorganic iodineCarpenter, Lucy J.; MacDonald, Samantha M.; Shaw, Marvin D.; Kumar, Ravi; Saunders, Russell W.; Parthipan, Rajendran; Wilson, Julie; Plane, John M. C.Nature Geoscience (2013), 6 (2), 108-111CODEN: NGAEBU; ISSN:1752-0894. (Nature Publishing Group)Naturally occurring bromine- and iodine-contg. compds. substantially reduce regional, and possibly even global, tropospheric ozone levels. As such, these halogen gases reduce the global warming effects of ozone in the troposphere, and its capacity to initiate the chem. removal of hydrocarbons such as methane. The majority of halogen-related surface ozone destruction is attributable to iodine chem. So far, org. iodine compds. have been assumed to serve as the main source of oceanic iodine emissions. However, known org. sources of atm. iodine cannot account for gas-phase iodine oxide concns. in the lower troposphere over the tropical oceans. Here, we quantify gaseous emissions of inorg. iodine following the reaction of iodide with ozone in a series of lab. expts. We show that the reaction of iodide with ozone leads to the formation of both mol. iodine and hypoiodous acid. Using a kinetic box model of the sea surface layer and a one-dimensional model of the marine boundary layer, we show that the reaction of ozone with iodide on the sea surface could account for around 75% of obsd. iodine oxide levels over the tropical Atlantic Ocean. According to the sea surface model, hypoiodous acid-not previously considered as an oceanic source of iodine-is emitted at a rate ten-fold higher than that of mol. iodine under ambient conditions.
- 53Sofiev, M.; Soares, J.; Prank, M.; de Leeuw, G.; Kukkonen, J. A regional-to-global model of emission and transport of sea salt particles in the atmosphere. Journal of Geophysical Research (Atmospheres) 2011, 116, 21302, DOI: 10.1029/2010JD014713There is no corresponding record for this reference.
- 54Gantt, B.; Meskhidze, N.; Facchini, M. C.; Rinaldi, M.; Ceburnis, D.; O'Dowd, C. D. Wind speed dependent size-resolved parameterization for the organic mass fraction of sea spray aerosol. Atmospheric Chemistry and Physics - ATMOS CHEM PHYS 2011, 11, 8777– 8790, DOI: 10.5194/acp-11-8777-2011There is no corresponding record for this reference.
- 55Olenius, T.; Roldin, P. Role of gas-molecular cluster-aerosol dynamics in atmospheric new-particle formation. Sci. Rep. 2022, 12, 10135, DOI: 10.1038/s41598-022-14525-yThere is no corresponding record for this reference.
- 56Ning, A.; Liu, L.; Zhang, S.; Yu, F.; Du, L.; Ge, M.; Zhang, X. The critical role of dimethylamine in the rapid formation of iodic acid particles in marine areas. npj Climate and Atmospheric Science 2022, 5, 92, DOI: 10.1038/s41612-022-00316-9There is no corresponding record for this reference.
- 57Besel, V.; Kubecka, J.; Kurtén, T.; Vehkamäki, H. Impact of Quantum Chemistry Parameter Choices and Cluster Distribution Model Settings on Modeled Atmospheric Particle Formation Rates. J. Phys. Chem. A 2020, 124, 5931, DOI: 10.1021/acs.jpca.0c0398457https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXht1alsLvK&md5=8e749fba9da93b18694e415c1c060662Impact of Quantum Chemistry Parameter Choices and Cluster Distribution Model Settings on Modeled Atmospheric Particle Formation RatesBesel, Vitus; Kubecka, Jakub; Kurten, Theo; Vehkamaki, HannaJournal of Physical Chemistry A (2020), 124 (28), 5931-5943CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Various parameters effect on the rate of new particle formation predicted for a H2SO4-NH3 system using quantum chem. and cluster distribution dynamics simulations, here, Atm. Cluster Dynamics Code, was tested. Consistent consideration of the rotational symmetry no. of monomers (H2SO4 and NH3 mols., bisulfate ion and NH4+) led to a significant rise in predicted particle formation rate; including the rotational symmetry no. of clusters only slightly changed results, and only for conditions where charged clusters dominate particle formation rate. This was because most of clusters stable enough to participate in new particle formation have a rotational symmetry no. of 1; few exceptions to this rule are pos. charged clusters. Applying quasi-harmonic correction for low frequency vibrational modes tended to generally decrease predicted new particle formation rates and significantly altered the slope of the formation rate curve plotted against H2SO4 concn., a typical convention in atm. aerosol science. Cluster max. size effect explicitly included in simulations depended on simulation conditions. Errors arising from a limited set of clusters were higher for higher evapn. rates and tended increase with temp. Errors tended to be higher for lower vapor concns. Boundary conditions for out-growing clusters (counted as formed particles) had only a small effect on results, provided the definition was chem. reasonable and the set of simulated clusters was sufficiently large. A comparison of data from cosmics leaving outdoor droplets (CLOUD) chamber and a cluster distribution dynamics model using older quantum chem. input data showed improved agreement when using new input data and the proposed combination of symmetry/quasi-harmonic corrections.
- 58Myllys, N.; Kubečka, J.; Besel, V.; Alfaouri, D.; Olenius, T.; Smith, J. N.; Passananti, M. Role of base strength, cluster structure and charge in sulfuric-acid-driven particle formation. Atmospheric Chemistry and Physics 2019, 19, 9753– 9768, DOI: 10.5194/acp-19-9753-2019There is no corresponding record for this reference.
- 59Zhang, R.; Xie, H.-B.; Ma, F.; Chen, J.; Iyer, S.; Simon, M.; Heinritzi, M.; Shen, J.; Tham, Y. J.; Kurtén, T.; Worsnop, D. R.; Kirkby, J.; Curtius, J.; Sipilä, M.; Kulmala, M.; He, X.-C. Critical Role of Iodous Acid in Neutral Iodine Oxoacid Nucleation. Environ. Sci. Technol. 2022, 56, 14166– 14177, DOI: 10.1021/acs.est.2c0432859https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB38XisVajsLrM&md5=529b05e9e166104af0dc3f36a5a4445fCritical Role of Iodous Acid in Neutral Iodine Oxoacid NucleationZhang, Rongjie; Xie, Hong-Bin; Ma, Fangfang; Chen, Jingwen; Iyer, Siddharth; Simon, Mario; Heinritzi, Martin; Shen, Jiali; Tham, Yee Jun; Kurten, Theo; Worsnop, Douglas R.; Kirkby, Jasper; Curtius, Joachim; Sipila, Mikko; Kulmala, Markku; He, Xu-ChengEnvironmental Science & Technology (2022), 56 (19), 14166-14177CODEN: ESTHAG; ISSN:1520-5851. (American Chemical Society)Nucleation of neutral iodine particles has recently been found to involve both iodic acid (HIO3) and iodous acid (HIO2). However, the precise role of HIO2 in iodine oxoacid nucleation remains unclear. Herein, we probe such a role by investigating the cluster formation mechanisms and kinetics of (HIO3)m(HIO2)n (m = 0-4, n = 0-4) clusters with quantum chem. calcns. and atm. cluster dynamics modeling. When compared with HIO3, we find that HIO2 binds more strongly with HIO3 and also more strongly with HIO2. After accounting for ambient vapor concns., the fastest nucleation rate is predicted for mixed HIO3-HIO2 clusters rather than for pure HIO3 or HIO2 ones. Our calcns. reveal that the strong binding results from HIO2 exhibiting a base behavior (accepting a proton from HIO3) and forming stronger halogen bonds. Moreover, the binding energies of (HIO3)m(HIO2)n clusters show a far more tolerant choice of growth paths when compared with the strict stoichiometry required for sulfuric acid-base nucleation. Our predicted cluster formation rates and dimer concns. are acceptably consistent with those measured by the Cosmic Leaving Outdoor Droplets (CLOUD) expt. This study suggests that HIO2 could facilitate the nucleation of other acids beyond HIO3 in regions where base vapors such as ammonia or amines are scarce.
- 60Schmitz, G.; Elm, J. Assessment of the DLPNO Binding Energies of Strongly Noncovalent Bonded Atmospheric Molecular Clusters. ACS Omega 2020, 5, 7601– 7612, DOI: 10.1021/acsomega.0c0043660https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXlt1yktL0%253D&md5=865ba10b3e9832494d960adb64b36715Assessment of the DLPNO Binding Energies of Strongly Noncovalent Bonded Atmospheric Molecular ClustersSchmitz, Gunnar; Elm, JonasACS Omega (2020), 5 (13), 7601-7612CODEN: ACSODF; ISSN:2470-1343. (American Chemical Society)This work assessed the performance of DLPNO-CCSD(T0), DLPNO-MP2, and d. functional theory methods in calcg. binding energies of a representative test set of 45 atm. acid/acid, acid/base, and acid/water dimer clusters. Approx. method performance was compared to high level, explicitly correlated CCSD(F12*)(T)/complete basis set (CBS) ref. calcns. Of the tested d. functionals, ωB97X-D3(BJ) had the best performance with a mean deviation of 0.09 kcal/mol and a max. deviation of 0.83 kcal/mol. The RI-CC2/aug-cc-pV(T+d)Z level of theory severely over-predicted cluster binding energies with a mean deviation of -1.31 kcal/mol and a max. deviation up to -3.00 kcal/mol. Thus, RI-CC2/aug-cc-pV(T+d)Z should not be used to study atm. mol. clusters. DLPNO variants were tested with/without the inclusion of explicit correlation (F12) in the wave function, with different pair natural orbital (PNO) settings (loosePNO, normalPNO, tightPNO) and using double and triple zeta basis sets. Performance of DLPNO-MP2 methods was independent of PNO settings and yielded low mean deviations (≤-0.84 kcal/mol). However, DLPNO-MP2 required explicitly correlated wave functions to yield max. deviations <1.40 kcal/mol. To obtain high accuracy with max. deviation <∼1.0 kcal/mol, DLPNO-CCSD(T0)/aug-cc-pVTZ (normalPNO) calcns. or DLPNO-CCSD(T0)-F12/cc-pVTZ-F12 (normalPNO) calcns. were required. The most accurate level of theory was DLPNO-CCSD(T0)-F12/cc-pVTZ-F12 using a tightPNO criterion which yielded a mean deviation of 0.10 kcal/mol, with a max. deviation of 0.20 kcal/mol, vs. the CCSD(F12*)(T)/CBS ref.
- 61Elm, J.; Kristensen, K. Basis Set Convergence of the Binding Energies of Strongly Hydrogen-Bonded Atmospheric Clusters. Phys. Chem. Chem. Phys. 2017, 19, 1122– 1133, DOI: 10.1039/C6CP06851K61https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XhvFKjtb%252FO&md5=c44dbe667e0a1605a94c77283fbf5b08Basis set convergence of the binding energies of strongly hydrogen-bonded atmospheric clustersElm, Jonas; Kristensen, KasperPhysical Chemistry Chemical Physics (2017), 19 (2), 1122-1133CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)We investigate the basis set convergence of second order Moller Plesset perturbation theory for the binding energies of 11 strongly hydrogen-bonded clusters relevant to the atm. The binding energies are calcd. both as the uncorrected (UC) value, as well as by employing the counterpoise (CP) and the Same No. Of Optimized Parameters (SNOOP) correction schemes, with and without explicit correlation (F12). We find that the use of the F12 corrections is of utter importance for obtaining converged binding energies. Without F12 corrections at least a quadruple-ζ basis set is generally required to yield errors below 1 kcal mol-1 compared to the complete basis set (CBS) limit. Using coupled cluster methods we obtain a best est. of the CCSD(T) CBS limit of the binding energies of the considered clusters and compare it with approx. DLPNO-CCSD(T) and DFT methods. We identify a pragmatic approach, relying solely on a series of double-ζ basis set calcns., for obtaining results in remarkably good agreement with our CBS est.
- 62Klamt, A. Conductor-like screening model for real solvents: a new approach to the quantitative calculation of solvation phenomena. J. Phys. Chem. 1995, 99, 2224– 2235, DOI: 10.1021/j100007a06262https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXjsFaisb0%253D&md5=74ff536ca20d353b9f8525d9f25c51d5Conductor-like Screening Model for Real Solvents: A New Approach to the Quantitative Calculation of Solvation PhenomenaKlamt, AndreasJournal of Physical Chemistry (1995), 99 (7), 2224-35CODEN: JPCHAX; ISSN:0022-3654. (American Chemical Society)Starting from the question of why dielec. continuum models give a fairly good description of mols. in water and some other solvents, a totally new approach for the calcn. of solvation phenomena is presented. It is based on the perfect, i.e., conductor-like, screening of the solute mol. and a quant. calcn. of the deviations from ideality appearing in real solvents. Thus, a new point of view to solvation phenomena is presented, which provides an alternative access to many questions of scientific and tech. importance. The whole theory is based on the results of MO continuum solvation models. A few representative solvents are considered, and the use of the theory is demonstrated by the calcn. of vapor pressures, surface tensions, and octanol/water partition coeffs.
- 63Klamt, A.; Jonas, V.; Bürger, T.; Lohrenz, J. C. W. Refinement and parametrization of COSMO-RS. J. Phys. Chem. A 1998, 102, 5074– 5085, DOI: 10.1021/jp980017s63https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXjs1Srs7Y%253D&md5=e9043dfbc9123e2a9b7e7a77216db911Refinement and Parametrization of COSMO-RSKlamt, Andreas; Jonas, Volker; Buerger, Thorsten; Lohrenz, John C. W.Journal of Physical Chemistry A (1998), 102 (26), 5074-5085CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The continuum solvation model COSMO and its extension beyond the dielec. approxn. (COSMO-RS) have been carefully parametrized in order to optimally reproduce 642 data points for a variety of properties, i.e., ΔG of hydration, vapor pressure, and the partition coeffs. for octanol/water, benzene/water, hexane/water, and di-Et ether/water. Two hundred seventeen small to medium sized neutral mols., covering most of the chem. functionality of the elements H, C, N, O, and Cl, have been considered. An overall accuracy of 0.4 (rms) kcal/mol for chem. potential differences, corresponding to a factor of 2 in the equil. consts. under consideration, has been achieved. This was using only a single radius and one dispersion const. per element and a total no. of eight COSMO-RS inherent parameters. Most of these parameters were close to their theor. est. The optimized cavity radii agreed well with the widely accepted rule of 120% of van der Waals radii. The whole parametrization was based upon d. functional calcns. using DMol/COSMO. As a result of this sound parametrization, we are now able to calc. almost any chem. equil. in liq./liq. and vapor/liq. systems up to an accuracy of a factor 2 without the need of any addnl. exptl. data for solutes or solvents. This opens a wide range of applications in phys. chem. and chem. engineering.
- 64Eckert, F.; Klamt, A. Fast solvent screening via quantum chemistry: COSMO-RS approach. AIChE J. 2002, 48, 369– 385, DOI: 10.1002/aic.69048022064https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD38XhsFSksLg%253D&md5=0f1996a34219de62fa735128a95020a1Fast solvent screening via quantum chemistry: COSMO-RS approachEckert, Frank; Klamt, AndreasAIChE Journal (2002), 48 (2), 369-385CODEN: AICEAC; ISSN:0001-1541. (American Institute of Chemical Engineers)Conductor-like Screening Model for Real Solvents (COSMO-RS), a general and fast methodol. for the a priori prediction of thermophys. data of liqs. is presented. It is based on cheap unimol. quantum chem. calcns., which, combined with exact statistical thermodn., provide the information necessary for the evaluation of mol. interactions in liqs. COSMO-RS is an alternative to structure interpolating group contribution methods. The method is independent of exptl. data and generally applicable. A methodol. comparison with group contribution methods is given. The applicability of the COSMO-RS method to the goal of solvent screening is demonstrated at various examples of vapor-liq.-, liq.-liq.-, solid-liq.-equil. and vapor-pressure predictions.
- 65BIOVIA COSMOtherm, Release 2021; Dassault Systèmes. http://www.3ds.com. 2021.There is no corresponding record for this reference.
- 66BIOVIA COSMOconf 2021; Dassault Systèmes. http://www.3ds.com. 2021.There is no corresponding record for this reference.
- 67Turbomole, V7.5.1; University of Karlsruhe and Forschungszentrum Karlsruhe GmbH, 2020.There is no corresponding record for this reference.
- 68Roldin, P.; Swietlicki, E.; Massling, A.; Kristensson, A.; Löndahl, J.; Eriksson, A.; Pagels, J.; Gustafsson, S. Aerosol ageing in an urban plume - implication for climate. Atmospheric Chemistry and Physics - ATMOS CHEM PHYS 2011, 11, 5897– 5915, DOI: 10.5194/acp-11-5897-2011There is no corresponding record for this reference.
- 69Jacobson, M. Z. Fundamentals of Atmospheric Modeling, 2nd ed.; Cambridge University Press, 2005.There is no corresponding record for this reference.
- 70BIOVIA COSMOtherm, version C3.0, Release 19; Dassault Systemes, 2019.There is no corresponding record for this reference.
- 71Pruppacher, H.; Klett, J.; Wang, P. Microphysics of Clouds and Precipitation. Aerosol Sci. Technol. 1998, 28, 381– 382, DOI: 10.1080/02786829808965531There is no corresponding record for this reference.
- 72Kurosaki, Y.; Matoba, S.; Iizuka, Y.; Fujita, K.; Shimada, R. Increased oceanic dimethyl sulfide emissions in areas of sea ice retreat inferred from a Greenland ice core. Communications Earth & Environment 2022, 3, 327, DOI: 10.1038/s43247-022-00661-wThere is no corresponding record for this reference.
Supporting Information
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.est.4c01891.
Figure S1 contains modeled and measured gas-phase concentrations of SO2, HIO3, and NOx, along with modeled gas-phase concentrations of OH, NH3, and DMA; Figure S2 illustrates the back trajectory heat maps for the full simulation period along with MP1, MP2, MP3, and CP1; Figure S3 demonstrates the method behind the air-mass origin analysis; Figure S4 contains the modeled size resolved chemical composition of SO4, NO3, Cl, NH4, Na, MSA, HIO3, DMA, SOA, and POA between 1 nm and 1 μm; Figure S5 illustrates the modeled and measured median particle number size distribution, including results from the woDMS, woIodine, and woAnthro sensitivity runs; Figure S6 contains modeled and measured particle numbers size distributions from Pallas, while Figure S7 contains similar results from Hyltemossa; Figure S8 illustrates the DMS, SO2, and OH gas-phase concentration along with the SA-NH3 nucleation rate along one of the HYSPLIT trajectories moving from the Norwegian Sea toward the SMEARII stationl Table S1 comprises COSMOtherm-derived Henry’s law solubilities, while Table S2 contains COSMOtherm-derived pKa values (PDF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.