Enhanced Organic Nitrate Formation from Peroxy Radicals in the Condensed PhaseClick to copy article linkArticle link copied!
- Victoria P. Barber*Victoria P. Barber*Email: [email protected]Department of Civil and Environmental Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, United StatesMore by Victoria P. Barber
- Lexy N. LeMarLexy N. LeMarDepartment of Chemical Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, United StatesMore by Lexy N. LeMar
- Yaowei LiYaowei LiJohn A. Paulson School of Engineering and Applied Sciences, Harvard University, Cambridge, Massachusetts 02138, United StatesMore by Yaowei Li
- Jonathan W. ZhengJonathan W. ZhengDepartment of Chemical Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, United StatesMore by Jonathan W. Zheng
- Frank N. KeutschFrank N. KeutschJohn A. Paulson School of Engineering and Applied Sciences, Harvard University, Cambridge, Massachusetts 02138, United StatesMore by Frank N. Keutsch
- Jesse H. Kroll*Jesse H. Kroll*Email: [email protected]Department of Civil and Environmental Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, United StatesDepartment of Chemical Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, United StatesMore by Jesse H. Kroll
Abstract
Organic alkoxy (RO) and peroxy (RO2) radicals are key intermediates in multiphase atmospheric oxidation chemistry, though most of the study of their chemistry has focused on the gas phase. To better understand how radical chemistry may vary across different phases, we examine the chemistry of a model system, the 1-pentoxy radical, in three phases: the aqueous phase, the condensed organic phase, and the gas phase. In each phase, we generate the 1-pentoxy radical from the photolysis of n-pentyl nitrite, run the chemistry under conditions in which RO2 radicals react with NO, and detect the products in real time using an ammonium chemical ionization mass spectrometer (NH4+ CIMS). The condensed-phase chemistry shows an increase in formation of organic nitrate (RONO2) from the downstream RO2+NO reaction, which is attributed to potential collisional and solvent-cage stabilization of the RO2–NO complex. We further observe an enhancement in the yield of carbonyl relative to hydroxy carbonyl products in the condensed phase, indicating changes to RO radical kinetics. The different branching ratios in the condensed phase impact the product volatility distribution as well as HOx-NOx chemistry, and may have implications for nitrate formation, aqueous aerosol formation, and radical cycling within atmospheric particles and droplets.
This publication is licensed under
License Summary*
You are free to share(copy and redistribute) this article in any medium or format within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
Non-Commercial (NC): Only non-commercial uses of the work are permitted.
No Derivatives (ND): Derivative works may be created for non-commercial purposes, but sharing is prohibited.
*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 within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
Non-Commercial (NC): Only non-commercial uses of the work are permitted.
No Derivatives (ND): Derivative works may be created for non-commercial purposes, but sharing is prohibited.
*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 within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
Non-Commercial (NC): Only non-commercial uses of the work are permitted.
No Derivatives (ND): Derivative works may be created for non-commercial purposes, but sharing is prohibited.
*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 within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
Non-Commercial (NC): Only non-commercial uses of the work are permitted.
No Derivatives (ND): Derivative works may be created for non-commercial purposes, but sharing is prohibited.
*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 within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
Non-Commercial (NC): Only non-commercial uses of the work are permitted.
No Derivatives (ND): Derivative works may be created for non-commercial purposes, but sharing is prohibited.
*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.
1. Introduction
2. Methods and Materials
3. Results and Discussion
Figure 1
Figure 1. : Time-resolved mass spectrometric signals for n-pentyl nitrite photolysis products, in three phases: (a) the gas phase, (b) the condensed organic phase, and (3) the aqueous phase. All measurements use the NH4+ CIMS, and formulas listed are those of the analytes (products are detected as the complexes with NH4+). Signals are smoothed over 15-s intervals and normalized to the dilution tracer, acetonitrile; t = 0 refers to the time at which UV lights are switched on.
Figure 2
Figure 2. : Simplified reaction mechanism for the photolysis of n-pentyl nitrite. Detected closed-shell products are highlighted, with colors corresponding to the products shown in Figure 1.
Photolysis rate constant | Hydrolysis rate constant | Alkoxy radical branching ratio | Nitrate branching ratio from RO2 + NO | |
---|---|---|---|---|
Phase | j1 (s–1) | k8 (s–1) | k3[O2]/(k3[O2] + k2) | X5a |
Gas | (2.9 ± 0.1) × 10–4 | N/A | (6.6 ± 0.4) × 10–3 | 0.22 ± 0.001 |
Organic | (4.3 ± 0.6) × 10–3 | N/A | 0.099 ± 0.008 | 0.86 ± 0.04 |
Aqueous | (7.0 ± 0.2) × 10–5 | (1.6 ± 0.7) × 10–3 | 0.31 ± 0.007 | 0.79 ± 0.03 |
Peroxy Radical | X5a(g), 1 atm, 298 K | X5a(aq), 298 K | X5a(aq)/X5a(g) |
---|---|---|---|
Methyl Peroxy | 0.017 ± 0.0014 (25) | 0.23 ± 0.4 (24) 0.86 ± 0.02 (28) | 14, 51 |
Ethyl Peroxy | 0.033 ± 0.004 (29) | 0.67 ± 0.03 (24) | 20 |
Propyl Peroxy | 0.038 ± 0.002 (30),a 0.036 ± 0.005 (31),b | 0.71 ± 0.04 (24),b | 19 |
Tert-Butanol Peroxy | 0.044 (23) | 0.86 ± 0.02 (28) | 20 |
5-hydroxy-pentan-2-yl peroxy | 0.13 (23) | 0.24 (23) | 1.8 (23) 3.5c |
All values are taken from measurements, apart from italicized items which are calculated from the structure–activity relationship presented by Jenkin et al. (23) Reported value for isopropyl peroxy radical only.
Reported value for a combination of n-propyl and isopropyl peroxy radicals.
This work.
4. Conclusions and Atmospheric Relevance
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.estlett.4c00473.
Additional details on experimental methods, including a diagram of the experimental setup; additional results, including an analysis of the pH dependence of n-pentyl nitrite hydrolysis, an analysis of tubing effects on the observed kinetics, and results of low-NO gas-phase experiments; a detailed derivation of the kinetic model used for extracting rate constants and branching ratios; a comparison between modeled and measured time profiles (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
This work is supported by NSF grant CHE-210881. The authors gratefully acknowledge William H. Green (MIT), and Hartmut Herrmann and Erik H. Hoffmann (Leibniz-Institute for Tropospheric Research) for helpful discussions.
References
This article references 47 other publications.
- 1Hoffmann, 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 (42), 11776– 11781, DOI: 10.1073/pnas.1606320113Google Scholar1An 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.
- 2George, I. J.; Abbatt, J. P. D. Heterogeneous Oxidation of Atmospheric Aerosol Particles by Gas-Phase Radicals. Nat. Chem. 2010, 2 (9), 713– 722, DOI: 10.1038/nchem.806Google Scholar2Heterogeneous oxidation of atmospheric aerosol particles by gas-phase radicalsGeorge, I. J.; Abbatt, J. P. D.Nature Chemistry (2010), 2 (9), 713-722CODEN: NCAHBB; ISSN:1755-4330. (Nature Publishing Group)A review concerning the heterogeneous oxidn. of org. and inorg. constituents of atm. aerosols by gas-phase radicals is given, focusing on the kinetics and reaction mechanisms and how they affect particle physicochem. properties (compn., size, d., hygroscopicity). Potential atm. impacts include: release of chem. reactive gases, e.g., halogens, aldehydes and org. acids; reactive loss of particle-borne mol. tracer and toxic species; and enhanced hygroscopic properties of aerosols which may improve their ability to form cloud droplets. Topics discussed include: radical uptake on atm. relevant surfaces; oxidn. of condensed-phase org. and inorg. compds.; particle physicochem. property modification; atm. implications; and outlook.
- 3Lambe, A. T.; Cappa, C. D.; Massoli, P.; Onasch, T. B.; Forestieri, S. D.; Martin, A. T.; Cummings, M. J.; Croasdale, D. R.; Brune, W. H.; Worsnop, D. R.; Davidovits, P. Relationship between Oxidation Level and Optical Properties of Secondary Organic Aerosol. Environ. Sci. Technol. 2013, 47 (12), 6349– 6357, DOI: 10.1021/es401043jGoogle Scholar3Relationship between Oxidation Level and Optical Properties of Secondary Organic AerosolLambe, Andrew T.; Cappa, Christopher D.; Massoli, Paola; Onasch, Timothy B.; Forestieri, Sara D.; Martin, Alexander T.; Cummings, Molly J.; Croasdale, David R.; Brune, William H.; Worsnop, Douglas R.; Davidovits, PaulEnvironmental Science & Technology (2013), 47 (12), 6349-6357CODEN: ESTHAG; ISSN:0013-936X. (American Chemical Society)Brown carbon (BrC), which may include secondary org. aerosol (SOA), can be a significant climate-forcing agent through its optical absorption properties; however, the overall contribution of SOA to BrC remains poorly understood. This work examd. correlations between oxidn. level and SOA optical properties. SOA was generated in a flow reactor in the absence of NOx by OH- oxidn. of gaseous precursors as surrogates for anthropogenic (naphthalene, tricyclo[5.2.1.02,6]decane), biomass burning (guaiacol), and biogenic (α-pinene) emissions. SOA chem. compn. was characterized by time-of-flight aerosol mass spectrometry. SOA mass-specific absorption cross sections (MAC) and refractive indexes were calcd. from real-time cavity ring-down photo-acoustic spectrometry measurements at 405 and 532 nm and from UV-vis spectrometry measurements of methanol exts. of filter-collected particles (300-600 nm). At 405 nm, SOA MAC values and imaginary refractive indexes increased with increasing oxidn. level and decreased with increasing wavelength, leading to negligible absorption at 532 nm. SOA real refractive indexes decreased with increasing oxidn. level. A literature study comparison suggested that under typical polluted conditions, the effect of NOx on SOA absorption was small. SOA may significantly contribute to atm. BrC; the magnitude depends on precursor type and oxidn. level.
- 4Galeazzo, T.; Valorso, R.; Li, Y.; Camredon, M.; Aumont, B.; Shiraiwa, M. Estimation of Secondary Organic Aerosol Viscosity from Explicit Modeling of Gas-Phase Oxidation of Isoprene and α-Pinene. Atmospheric Chem. Phys. 2021, 21 (13), 10199– 10213, DOI: 10.5194/acp-21-10199-2021Google Scholar4Estimation of secondary organic aerosol viscosity from explicit modeling of gas-phase oxidation of isoprene and α-pineneGaleazzo, Tommaso; Valorso, Richard; Li, Ying; Camredon, Marie; Aumont, Bernard; Shiraiwa, ManabuAtmospheric Chemistry and Physics (2021), 21 (13), 10199-10213CODEN: ACPTCE; ISSN:1680-7324. (Copernicus Publications)Secondary org. aerosols (SOA) are major components of atm. fine particulate matter, affecting climate and air quality. Mounting evidence exists that SOA can adopt glassy and viscous semisolid states, impacting formation and partitioning of SOA. In this study, we apply the GECKO-A (Generator of Explicit Chem. and Kinetics of Orgs. in the Atm.) model to conduct explicit chem. modeling of isoprene photooxidn. and α-pinene ozonolysis and their subsequent SOA formation. The detailed gas-phase chem. schemes from GECKO-A are implemented into a box model and coupled to our recently developed glass transition temp. parameterizations, allowing us to predict SOA viscosity. The effects of chem. compn., relative humidity, mass loadings and mass accommodation on particle viscosity are investigated in comparison with measurements of SOA viscosity. The simulated viscosity of isoprene SOA agrees well with viscosity measurements as a function of relative humidity, while the model underestimates viscosity of α-pinene SOA by a few orders of magnitude. This difference may be due to missing processes in the model, including autoxidn. and particle-phase reactions, leading to the formation of high-molar-mass compds. that would increase particle viscosity. Addnl. simulations imply that kinetic limitations of bulk diffusion and redn. in mass accommodation coeff. may play a role in enhancing particle viscosity by suppressing condensation of semi-volatile compds. The developed model is a useful tool for anal. and investigation of the interplay among gas-phase reactions, particle chem. compn. and SOA phase state.
- 5Kroll, J. H.; Seinfeld, J. H. Chemistry of Secondary Organic Aerosol: Formation and Evolution of Low-Volatility Organics in the Atmosphere. Atmos. Environ. 2008, 42 (16), 3593– 3624, DOI: 10.1016/j.atmosenv.2008.01.003Google Scholar5Chemistry of secondary organic aerosol: Formation and evolution of low-volatility organics in the atmosphereKroll, Jesse H.; Seinfeld, John H.Atmospheric Environment (2008), 42 (16), 3593-3624CODEN: AENVEQ; ISSN:1352-2310. (Elsevier Ltd.)A review is given. Secondary org. aerosol (SOA), particulate matter composed of compds. formed from the atm. transformation of org. species, accounts for a substantial fraction of tropospheric aerosol. The formation of low-volatility (semivolatile and possibly nonvolatile) compds. that make up SOA is governed by a complex series of reactions of a large no. of org. species, so the exptl. characterization and theor. description of SOA formation presents a substantial challenge. We outline what is known about the chem. of formation and continuing transformation of low-volatility species in the atm. The primary focus is chem. processes that can change the volatility of org. compds.: (1) oxidn. reactions in the gas phase, (2) reactions in the particle phase, and (3) continuing chem. (in either phase) over several generations. Gas-phase oxidn. reactions can reduce volatility by the addn. of polar functional groups or increase it by the cleavage of carbon-carbon bonds; key branch points that control volatility are the initial attack of the oxidant, reactions of alkylperoxy (RO2) radicals, and reactions of alkoxy (RO) radicals. Reactions in the particle phase include oxidn. reactions as well as accretion reactions, non-oxidative processes leading to the formation of high-mol.-wt. species. Org. C in the atm. is continually subject to reactions in the gas and particle phases throughout its atm. lifetime (until lost by phys. deposition or oxidized to CO or CO2), implying continual changes in volatility over the timescales of several days. The volatility changes arising from these chem. reactions must be parameterized and included in models in order to gain a quant. and predictive understanding of SOA formation.
- 6Calvert, J. G.; Derwent, R. G.; Orlando, J. J.; Wallington, T. J.; Tyndall, G. S. Mechanisms of Atmospheric Oxidation of the Alkanes; Oxford University Press: Oxford, NY, 2008.Google ScholarThere is no corresponding record for this reference.
- 7Atkinson, R.; Arey, J. Atmospheric Degradation of Volatile Organic Compounds. Chem. Rev. 2003, 103 (12), 4605– 4638, DOI: 10.1021/cr0206420Google Scholar7Atmospheric Degradation of Volatile Organic CompoundsAtkinson, Roger; Arey, JanetChemical Reviews (Washington, DC, United States) (2003), 103 (12), 4605-4638CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review concerning the removal of biogenic and anthropogenic volatile org. compds. (VOC) emitted to the atm. through phys. (deposition) or chem. processes, or their atm. transformation is given. Topics discussed include: tropospheric VOC transformation processes (initial reactions and lifetimes, reaction mechanisms); atm. chem. of alkanes (kinetic data for initial OH- and NO3- reactions, reaction mechanism); atm. chem. of alkenes (rate consts. for initial reaction of alkenes with OH-, NO3-, and O3; mechanism of the OH- reaction; NO3- reaction; reaction with O3); arom. hydrocarbons (kinetics of OH- reactions, reactions of phenols, reactions of unsatd. 1,4-dicarbonyls and di-unsatd. 1,6-dicarbonyls); atm. reactions of oxygenated VOC (aldehydes, ketones, aliph. alcs., ethers, alkyl nitrates); and conclusions and future research needs (exptl. studies, theor. studies and crit. reviews and evaluations).
- 8Calvert, J. G.; Mellouki, A.; Orlando, J. J. Mechanisms of Atmospheric Oxidation of the Oxygenates; Oxford University Press: Oxford, NY, 2011. DOI: 10.1093/oso/9780195365818.005.0001 .Google ScholarThere is no corresponding record for this reference.
- 9Leviss, D. H.; Van Ry, D. A.; Hinrichs, R. Z. Multiphase Ozonolysis of Aqueous α-Terpineol. Environ. Sci. Technol. 2016, 50 (21), 11698– 11705, DOI: 10.1021/acs.est.6b03612Google Scholar9Multiphase Ozonolysis of Aqueous α-TerpineolLeviss, Dani H.; Van Ry, Daryl A.; Hinrichs, Ryan Z.Environmental Science & Technology (2016), 50 (21), 11698-11705CODEN: ESTHAG; ISSN:0013-936X. (American Chemical Society)Multiphase ozonolysis of aq. orgs. presents a potential pathway for the formation of aq. secondary org. aerosol (aqSOA). We investigated the multiphase ozonolysis of α-terpineol, an oxygenated deriv. of limonene, and found that the reaction products and kinetics differ from the gas-phase ozonolysis of α-terpineol. One- and two-dimensional NMR spectroscopies along with GC-MS identified the aq. ozonolysis reaction products as trans- and cis-lactols [4-(5-hydroxy-2,2-dimethyltetrahydrofuran-3-yl)butan-2-one] and a lactone [4-hydroxy-4-methyl-3-(3-oxobutyl)-valeric acid gamma-lactone], which accounted for 46%, 27%, and 20% of the obsd. products, resp. Hydrogen peroxide was also formed in 10% yield consistent with a mechanism involving decompn. of hydroxyl hydroperoxide intermediates followed by hemiacetal ring closure. Multiphase reaction kinetics at gaseous ozone concns. of 131, 480, and 965 parts-per-billion were analyzed using a resistance model of net ozone uptake and found the second-order rate coeff. for the aq. reaction of α-terpineol + O3 to be 9.9(±3.3) × 106 M-1 s-1. Multiphase ozonolysis will therefore be competitive with multiphase oxidn. by hydroxyl radicals (OH) and ozonolysis of gaseous α-terpineol. We also measured product yields for the heterogeneous ozonolysis of α-terpineol adsorbed on glass, NaCl, and kaolinite, and identified the same three major products but with an increasing lactone yield of 33, 49, and 55% on these substrates, resp.
- 10Akimoto, H.; Hirokawa, J. Atmospheric Multiphase Chemistry: Fundamentals of Secondary Aerosol Formation; John Wiley & Sons, 2020. DOI: 10.1002/9781119422419 .Google ScholarThere is no corresponding record for this reference.
- 11Thornberry, T.; Abbatt, J. P. D. Heterogeneous Reaction of Ozone with Liquid Unsaturated Fatty Acids: Detailed Kinetics and Gas-Phase Product Studies. Phys. Chem. Chem. Phys. 2004, 6 (1), 84– 93, DOI: 10.1039/b310149eGoogle Scholar11Heterogeneous reaction of ozone with liquid unsaturated fatty acids: detailed kinetics and gas-phase product studiesThornberry, T.; Abbatt, J. P. D.Physical Chemistry Chemical Physics (2004), 6 (1), 84-93CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)Detailed kinetic and product yield studies have been performed for the heterogeneous reaction between gas-phase ozone and three liq. fatty acids using a coated-wall flow tube and chem. ionization mass spectrometry. Gas-surface reaction probabilities for ozone loss of (8.0 ±1.0) × 10-4, (1.3 ±0.1) × 10-3, and (1.8 ±0.2) × 10-3 have been measured at room temp. (298 K) for oleic acid, linoleic acid, and linolenic acid, resp. The temp. dependence of the uptake coeffs. was found to be small and pos. Comparison of these results to the kinetics of the equiv. gas-phase reactions implies that there is a definite enhancement in the rate for the heterogeneous process due to entropic factors, i.e., due to collisional trapping of ozone in the surface layers of the liq., and a possible effect on the activation energy of the reaction. For linoleic acid, the reaction probability was found to be independent of relative humidity (up to 55%), to ±10%, at 263 K. Volatile reaction products were obsd. using proton-transfer-reaction mass spectrometry. Nonanal was obsd. with a 0.50 (±0.10) yield for the reaction with oleic acid, whereas hexanal and nonenal were obsd. for linoleic acid with 0.25 (±0.05) and 0.29 (±0.05) yields, resp. These results indicate that the primary ozonide formed initially in the reaction can decomp. via two equal probability pathways and that a secondary ozonide is not formed in high yield in the aldehydic channel. These reactions represent a source of oxygenates to the atm. and will modify the hygroscopic properties of aerosols.
- 12Adams, G. E.; Willson, R. L. Pulse Radiolysis Studies on the Oxidation of Organic Radicals in Aqueous Solution. Trans. Faraday Soc. 1969, 65, 2981, DOI: 10.1039/tf9696502981Google Scholar12Pulse radiolysis studies on the oxidation of organic radicals in aqueous solutionAdams, Gerald Edward; Willson, Robin LindhopeTransactions of the Faraday Society (1969), 65 (11), 2981-7CODEN: TFSOA4; ISSN:0014-7672.Pulse radiolysis has been used to measure directly the abs. rates of oxidn. by ferricyanide ion of various radicals produced by OH attack on org. solutes. These include mono-, di-, and polyhydroxylic compds., hydroxy acids, polyethylene oxides of mol. wt. 200, 6000, and 20,000 and the amino acid serine. Radicals produced by H abstraction from α-C atoms in alcs. are oxidized at, or near, diffusion-controlled rates, whereas the reactions are much slower for radicals formed by OH-attack elsewhere. The technique has been used to measure the percentage of OH attack at the α-position for a series of straight and branched-chain alcs. O competes with ferricyanide for radical oxidn. The data for O-contg. solns. fit a simple radical-competition scheme which has been used to measure rates of peroxy-radical formation. These approach diffusion-controlled limits.
- 13Schuchmann, M. N.; Von Sonntag, C. The Rapid Hydration of the Acetyl Radical. A Pulse Radiolysis Study of Acetaldehyde in Aqueous Solution. J. Am. Chem. Soc. 1988, 110 (17), 5698– 5701, DOI: 10.1021/ja00225a019Google Scholar13The rapid hydration of the acetyl radical. A pulse radiolysis study of acetaldehyde in aqueous solution.Schuchmann, Man Nien; Von Sonntag, ClemensJournal of the American Chemical Society (1988), 110 (17), 5698-701CODEN: JACSAT; ISSN:0002-7863.In aq. soln. acetaldehyde and its hydrate are in a 0.8:1 equil. Hydroxyl radicals, generated by the pulse radiolysis of N2O-satd. water, react with acetaldehyde about 3 times faster than with the hydrate. The predominant radicals formed are the acetyl radical and its hydrated form, H-abstraction at Me occurring to only about 5-10%. The acetyl radical rapidly hydrates. Its rate of hydration is 2 × 106 times faster than that of acetaldehyde. The hydrated acetyl radical has been monitored by its rapid redn. of tetranitromethane, the formation of O2- in the presence of Oxygen, and its deprotonation at pH 11. The acetyl radical does not oxidize N,N,N',N'-tetramethyl-p-phenylenediamine (TMPD) on the pulse radiolysis time scale. The acetylperoxyl radical (formed in the presence of oxygen) reacts rapidly with TMPD, ascorbate, and O2-; it is the most strongly oxidizing peroxyl radical known so far. Data on the rate of water loss from the hydrated acetyl radical are considerably less accurate, but the rate const. in estd.
- 14Mezyk, S. P.; Elliot, A. J. Pulse Radiolysis of Iodate in Aqueous Solution. J. Chem. Soc. Faraday Trans. 1994, 90 (6), 831– 836, DOI: 10.1039/ft9949000831Google ScholarThere is no corresponding record for this reference.
- 15Lee, A. K. Y.; Zhao, R.; Gao, S. S.; Abbatt, J. P. D. Aqueous-Phase OH Oxidation of Glyoxal: Application of a Novel Analytical Approach Employing Aerosol Mass Spectrometry and Complementary Off-Line Techniques. J. Phys. Chem. A 2011, 115 (38), 10517– 10526, DOI: 10.1021/jp204099gGoogle Scholar15Aqueous-phase OH oxidation of glyoxal: application of a novel analytical approach employing aerosol mass spectrometry and complementary off-line techniquesLee, Alex K. Y.; Zhao, R.; Gao, S. S.; Abbatt, J. P. D.Journal of Physical Chemistry A (2011), 115 (38), 10517-10526CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Aq.-phase chem. of glyoxal may play an important role in the formation of highly oxidized secondary org. aerosol (SOA) in the atm. In this work, we use a novel design of photochem. reactor that allows for simultaneous photo-oxidn. and atomization of a bulk soln. to study the aq.-phase OH oxidn. of glyoxal. By employing both online aerosol mass spectrometry (AMS) and offline ion chromatog. (IC) measurements, glyoxal and some major products including formic acid, glyoxylic acid, and oxalic acid in the reacting soln. were simultaneously quantified. This is the first attempt to use AMS in kinetics studies of this type. The results illustrate the formation of highly oxidized products that likely coexist with traditional SOA materials, thus, potentially improving model predictions of org. aerosol mass loading and degree of oxidn. Formic acid is the major volatile species identified, but the atm. relevance of its formation chem. needs to be further investigated. While successfully quantifying low mol. wt. org. oxygenates and tentatively identifying a reaction product formed directly from glyoxal and hydrogen peroxide, comparison of the results to the offline total org. carbon (TOC) anal. clearly shows that the AMS is not able to quant. monitor all dissolved orgs. in the bulk soln. This is likely due to their high volatility or low stability in the evapd. soln. droplets. This exptl. approach simulates atm. aq. phase processing by conducting oxidn. in the bulk phase, followed by evapn. of water and volatile orgs. to form SOA.
- 16Carrasquillo, A. J.; Hunter, J. F.; Daumit, K. E.; Kroll, J. H. Secondary Organic Aerosol Formation via the Isolation of Individual Reactive Intermediates: Role of Alkoxy Radical Structure. J. Phys. Chem. A 2014, 118 (38), 8807– 8816, DOI: 10.1021/jp506562rGoogle Scholar16Secondary Organic Aerosol Formation via the Isolation of Individual Reactive Intermediates: Role of Alkoxy Radical StructureCarrasquillo, Anthony J.; Hunter, James F.; Daumit, Kelly E.; Kroll, Jesse H.Journal of Physical Chemistry A (2014), 118 (38), 8807-8816CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The study of the chem. underlying secondary org. aerosol (SOA) formation is complicated by the large no. of reaction pathways and oxidn. generations available to a given precursor species. Here we simplify such complexity to that of a single alkoxy radical (RO), by forming SOA via the direct photolysis of alkyl nitrite (RONO) isomers. Chamber expts. were conducted with 11 C10 RONO isomers to det. how the position of the radical center and branching of the carbon skeleton influences SOA formation. SOA yields served as a probe of RO reactivity, with lower yields indicating that fragmentation reactions dominate and higher yields suggesting the predominance of RO isomerization. The largest yields were from straight-chain isomers, particularly those with radical centers located toward the terminus of the mol. Trends in SOA yields can be explained in terms of two major effects: (1) the relative importance of isomerization and fragmentation reactions, which control the distribution of products, and (2) differences in volatility among the various isomeric products formed. Yields from branched isomers, which were low but variable, provide insight into the degree of fragmentation of the alkoxy radicals; in the case of the two β-substituted alkoxy radicals, fragmentation appears to occur to a greater extent than predicted by structure-activity relationships. Our results highlight how subtle differences in alkoxy radical structure can have major impacts on product yields and SOA formation.
- 17Kessler, S. H.; Nah, T.; Carrasquillo, A. J.; Jayne, J. T.; Worsnop, D. R.; Wilson, K. R.; Kroll, J. H. Formation of Secondary Organic Aerosol from the Direct Photolytic Generation of Organic Radicals. J. Phys. Chem. Lett. 2011, 2 (11), 1295– 1300, DOI: 10.1021/jz200432nGoogle Scholar17Formation of Secondary Organic Aerosol from the Direct Photolytic Generation of Organic RadicalsKessler, Sean H.; Nah, Theodora; Carrasquillo, Anthony J.; Jayne, John T.; Worsnop, Douglas R.; Wilson, Kevin R.; Kroll, Jesse H.Journal of Physical Chemistry Letters (2011), 2 (11), 1295-1300CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)The immense complexity inherent in secondary org. aerosol (SOA) formation, primarily due to the large no. of oxidn. steps and reaction pathways involved, has limited a detailed understanding of its underlying chem. To simplify such complexity, this work demonstrates SOA formation via photolysis of gas-phase alkyl iodides, generating org. peroxy radicals with known structures. In contrast to std. OH--initiated oxidn. expts., photolytically-initiated oxidn. forms a limited no. of products in a single reactive step. Typical for SOA, yields of aerosol generated from alkyl iodide photolysis depends on aerosol load, indicating the semi-volatile nature of the particulate species. However, aerosols were obsd. to have higher volatility and be less oxidized than previous multi-generational studies of alkane oxidn., suggesting addnl. oxidative steps are necessary to produce oxidized semi-volatile matter in the atm. Despite the relative simplicity of this chem. system, SOA mass spectra are still quite complex, underscoring the wide range of products present in SOA.
- 18Iglesias, E.; Casado, J. Mechanisms of Hydrolysis and Nitrosation Reactions of Alkyl Nitrites in Various Media. Int. Rev. Phys. Chem. 2002, 21 (1), 37– 74, DOI: 10.1080/01442350110092693Google ScholarThere is no corresponding record for this reference.
- 19Hunter, J. F.; Carrasquillo, A. J.; Daumit, K. E.; Kroll, J. H. Secondary Organic Aerosol Formation from Acyclic, Monocyclic, and Polycyclic Alkanes. Environ. Sci. Technol. 2014, 48 (17), 10227– 10234, DOI: 10.1021/es502674sGoogle Scholar19Secondary organic aerosol formation from acyclic, monocyclic, and polycyclic alkanesHunter, James F.; Carrasquillo, Anthony J.; Daumit, Kelly E.; Kroll, Jesse H.Environmental Science & Technology (2014), 48 (17), 10227-10234CODEN: ESTHAG; ISSN:0013-936X. (American Chemical Society)A large no. of org. species emitted into the atm. contain cycloalkyl groups. While cyclic species are believed to be important secondary org. aerosol (SOA) precursors, the specific role of cyclic moieties (particularly for species with multiple or fused rings) remains uncertain. Here we examine the yields and compn. of SOA formed from the reaction of OH with a series of C10 (cyclo)alkanes, with 0-3 rings, in order to better understand the role of multiple cyclic moieties on aerosol formation pathways. A chamber oxidn. technique using high, sustained OH radical concns. was used to simulate long reaction times in the atm. This aging technique leads to higher yields than in previously reported chamber expts. Yields were highest for cyclic and polycyclic precursors, though yield exhibited little dependence on no. of rings. However, the oxygen-to-carbon ratio of the SOA was highest for the polycyclic precursors. These trends are consistent with aerosol formation requiring two generations of oxidn. and 3-4 oxygen-contg. functional groups in order to condense. Cyclic, unbranched structures are protected from fragmentation during the first oxidn. step, with C-C bond scission instead leading to ring opening, efficient functionalization, and high SOA yields. Fragmentation may occur during subsequent oxidn. steps, limiting yields by forming volatile products. Polycyclic structures can undergo multiple ring opening reactions, but do not have markedly higher yields, likely due to enhanced fragmentation in the second oxidn. step. By contrast, C-C bond scission for the linear and branched structures leads to fragmentation prior to condensation, resulting in low SOA yields. The results highlight the key roles of multigenerational chem. and susceptibility to fragmentation in the formation and evolution of SOA.
- 20Zaytsev, A.; Breitenlechner, M.; Koss, A. R.; Lim, C. Y.; Rowe, J. C.; Kroll, J. H.; Keutsch, F. N. Using Collision-Induced Dissociation to Constrain Sensitivity of Ammonia Chemical Ionization Mass Spectrometry (NH4+ CIMS) to Oxygenated Volatile Organic Compounds. Atmospheric Meas. Technol. 2019, 12 (3), 1861– 1870, DOI: 10.5194/amt-12-1861-2019Google ScholarThere is no corresponding record for this reference.
- 21Orlando, J. J.; Tyndall, G. S.; Wallington, T. J. The Atmospheric Chemistry of Alkoxy Radicals. Chem. Rev. 2003, 103 (12), 4657– 4690, DOI: 10.1021/cr020527pGoogle Scholar21The Atmospheric Chemistry of Alkoxy RadicalsOrlando, John J.; Tyndall, Geoffrey S.; Wallington, Timothy J.Chemical Reviews (Washington, DC, United States) (2003), 103 (12), 4657-4689CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review concerning the atm. chem. of alkoxy radicals and rates and mechanisms of various reaction pathways is given. Topics discussed include: exptl. methods used to study alkoxy radical chem.; alkoxy radical chem. (reaction with O2 [CH3O• + O2; C2H5O• + O2; 1-C3H7O•, 2-C2H7O•, 1-C4H9O•, 2-C2H9O•, 3-C5H11O• + O2; CH2ClO•, CFCl2CH2O• + O2; comparison with previous recommendations], dissocn. reactions [unsubstituted alkoxy radicals, β-hydroxyalkoxy radicals, other O-substituted alkoxy radicals, halogenated alkoxy radicals], isomerization reactions, other intramol. reactions [HCl elimination reactions, the ester rearrangement reaction], chem. activation); and conclusions and suggestions for further study.
- 22Atkinson, R. Rate Constants for the Atmospheric Reactions of Alkoxy Radicals: An Updated Estimation Method. Atmos. Environ. 2007, 41 (38), 8468– 8485, DOI: 10.1016/j.atmosenv.2007.07.002Google Scholar22Rate constants for the atmospheric reactions of alkoxy radicals: An updated estimation methodAtkinson, RogerAtmospheric Environment (2007), 41 (38), 8468-8485CODEN: AENVEQ; ISSN:1352-2310. (Elsevier Ltd.)Alkoxy radicals are key intermediates in the atm. degrdns. of volatile org. compds., and can typically undergo reaction with O2, unimol. decompn. or unimol. isomerization. Previous structure-reactivity relationships for the estn. of rate consts. for these processes for alkoxy radicals have been updated to incorporate recent kinetic data from abs. and relative rate studies. Temp.-dependent rate expressions are derived allowing rate consts. for all three of these alkoxy radical reaction pathways to be calcd. at atmospherically relevant temps.
- 23Jenkin, M. E.; Valorso, R.; Aumont, B.; Rickard, A. R. Estimation of Rate Coefficients and Branching Ratios for Reactions of Organic Peroxy Radicals for Use in Automated Mechanism Construction. Atmospheric Chem. Phys. 2019, 19 (11), 7691– 7717, DOI: 10.5194/acp-19-7691-2019Google Scholar23Estimation of rate coefficients and branching ratios for reactions of organic peroxy radicals for use in automated mechanism constructionJenkin, Michael E.; Valorso, Richard; Aumont, Bernard; Rickard, Andrew R.Atmospheric Chemistry and Physics (2019), 19 (11), 7691-7717CODEN: ACPTCE; ISSN:1680-7324. (Copernicus Publications)Org. peroxy radicals (RO2), formed from the degrdn. of hydrocarbons and other volatile org. compds. (VOCs), play a key role in tropospheric oxidn. mechanisms. Several competing reactions may be available for a given RO2 radical, the relative rates of which depend on both the structure of RO2 and the ambient conditions. Published kinetics and branching ratio data are reviewed for the bimol. reactions of RO2 with NO, NO2, NO3, OH and HO2; and for their self-reactions and cross-reactions with other RO2 radicals. This information is used to define generic rate coeffs. and structure-activity relationship (SAR) methods that can be applied to the bimol. reactions of a series of important classes of hydrocarbon and oxygenated RO2 radicals. Information for selected unimol. isomerization reactions (i.e. H-atom shift and ring-closure reactions) is also summarized and discussed. The methods presented here are intended to guide the representation of RO2 radical chem. in the next generation of explicit detailed chem. mechanisms.
- 24Dahl, E. E.; Saltzman, E. S.; De Bruyn, W. J. The Aqueous Phase Yield of Alkyl Nitrates from ROO + NO: Implications for Photochemical Production in Seawater. Geophys. Res. Lett. 2003, 30 (6), 1271, DOI: 10.1029/2002GL016811Google ScholarThere is no corresponding record for this reference.
- 25Butkovskaya, N.; Kukui, A.; Le Bras, G. Pressure and Temperature Dependence of Methyl Nitrate Formation in the CH3O2 + NO Reaction. J. Phys. Chem. A 2012, 116 (24), 5972– 5980, DOI: 10.1021/jp210710dGoogle Scholar25Pressure and Temperature Dependence of Methyl Nitrate Formation in the CH3O2 + NO ReactionButkovskaya, Nadezhda; Kukui, Alexandre; Le Bras, GeorgesJournal of Physical Chemistry A (2012), 116 (24), 5972-5980CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The branching ratio β = k1b/k1a for the formation of Me nitrate, CH3ONO2, in the gas-phase CH3O2 + NO reaction, CH3O2 + NO → CH3O + NO2 (1a), CH3O2 + NO → CH3ONO2 (1b), has been detd. over the pressure and temp. ranges 50-500 Torr and 223-300 K, resp., using a turbulent flow reactor coupled with a chem. ionization mass spectrometer. At 298 K, the CH3ONO2 yield has been found to increase linearly with pressure from 0.33 ± 0.16% at 50 Torr to 0.80 ± 0.54% at 500 Torr (errors are 2σ). Decrease of temp. from 300 to 220 K leads to an increase of β by a factor of about 3 in the 100-200 Torr range. These data correspond to a value of β ≈ 1.0 ± 0.7% over the pressure and temp. ranges of the whole troposphere. Atm. concns. of CH3ONO2 roughly estd. using results of this work are in reasonable agreement with those obsd. in polluted environments and significantly higher compared with measurements in upper troposphere and lower stratosphere.
- 26Richards-Henderson, N. K.; Goldstein, A. H.; Wilson, K. R. Large Enhancement in the Heterogeneous Oxidation Rate of Organic Aerosols by Hydroxyl Radicals in the Presence of Nitric Oxide. J. Phys. Chem. Lett. 2015, 6 (22), 4451– 4455, DOI: 10.1021/acs.jpclett.5b02121Google Scholar26Large Enhancement in the Heterogeneous Oxidation Rate of Organic Aerosols by Hydroxyl Radicals in the Presence of Nitric OxideRichards-Henderson, Nicole K.; Goldstein, Allen H.; Wilson, Kevin R.Journal of Physical Chemistry Letters (2015), 6 (22), 4451-4455CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)In the troposphere, the heterogeneous lifetime of an org. mol. in an aerosol exposed to OH- is thought to be weeks, which is orders of magnitude slower than the analogous gas phase reactions (hours). This paper reports an unexpectedly large acceleration in the effective heterogeneous OH- reaction rate in the presence of NO. This 10-50 fold acceleration originates from free radical chain reactions, propagated by alkoxy radicals which form inside the aerosols by the NO reaction with peroxy radicals, which do not appear to produce chain-terminating products (e.g., alkyl nitrates), unlike gas phase mechanisms. A kinetic model, constrained by exptl. data, suggests that in polluted regions, heterogeneous oxidn. plays a much more prominent role in the daily chem. evolution of org. aerosols than previously believed.
- 27Renbaum, L. H.; Smith, G. D. Organic Nitrate Formation in the Radical-Initiated Oxidation of Model Aerosol Particles in the Presence of NOx. Phys. Chem. Chem. Phys. 2009, 11 (36), 8040– 8047, DOI: 10.1039/b909239kGoogle ScholarThere is no corresponding record for this reference.
- 28Goldstein, S.; Lind, J.; Merenyi, G. Reaction of Organic Peroxyl Radicals with •NO2 and •NO in Aqueous Solution: Intermediacy of Organic Peroxynitrate and Peroxynitrite Species. J. Phys. Chem. A 2004, 108 (10), 1719– 1725, DOI: 10.1021/jp037431zGoogle Scholar28Reaction of Organic Peroxyl Radicals with •NO2 and •NO in Aqueous Solution: Intermediacy of Organic Peroxynitrate and Peroxynitrite SpeciesGoldstein, Sara; Lind, Johan; Merenyi, GaborJournal of Physical Chemistry A (2004), 108 (10), 1719-1725CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)In this work, we studied the reactions of alkyl peroxyl radicals with •NO2 and •NO using the pulse radiolysis technique. The rate consts. for the reaction of •NO2 with (CH3)2C(OH)CH2OO•, CH3OO•, and c-C5H9OO• vary between 7 × 108 and 1.5 × 109 M-1 s-1. The reaction produces relatively long-lived alkyl peroxynitrates, which are in equil. with the parent radicals and have no appreciable absorption above 270 nm. It is also shown that •NO adds rapidly to (CH3)2C(OH)CH2OO• and CH3OO• to form alkyl peroxynitrites. The rate consts. for these reactions were detd. to be 2.8 × 109 and 3.5 × 109 M-1 s-1, resp. However, in contrast to alkyl peroxynitrates, alkyl peroxynitrites do not accumulate. Rather, they decomp. rapidly via homolysis along the relatively weak O-O bond, initially forming a geminate pair. Most of this pair collapses in the cage to form an alkyl nitrate, RONO2, and about 14% diffuses out as free alkoxyl and •NO2 radicals. A thermokinetic anal. predicts the half-life of CH3OONO in water to be less than 1 μs, an est. that agrees well with previous exptl. findings of ours for other alkyl peroxynitrites. A comparison of aq. and gaseous thermochem. of alkyl peroxynitrates reveals that alkyl peroxyl radicals and the corresponding alkyl peroxynitrates are similarly solvated by water.
- 29Butkovskaya, N.; Kukui, A.; Le Bras, G. Pressure and Temperature Dependence of Ethyl Nitrate Formation in the C2H5O2 + NO Reaction. J. Phys. Chem. A 2010, 114 (2), 956– 964, DOI: 10.1021/jp910003aGoogle Scholar29Pressure and Temperature Dependence of Ethyl Nitrate Formation in the C2H5O2 + NO ReactionButkovskaya, Nadezhda; Kukui, Alexandre; Le Bras, GeorgesJournal of Physical Chemistry A (2010), 114 (2), 956-964CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The branching ratio β = k1b/k1a for the formation of Et nitrate, C2H5ONO2, in the gas-phase C2H5O2 + NO reaction, C2H5O2 + NO → C2H5O + NO2 (1a), C2H5O2 + NO → C2H5ONO2 (1b), was detd. over the pressure and temp. ranges 100-600 Torr and 223-298 K, resp., using a turbulent flow reactor (TFR)-ion-mol. reactor(IMR) coupled with a chem. ionization mass spectrometer. At 298 K the C2H5ONO2 yield was found to increase linearly with pressure from about 0.7% at 100 Torr to about 3% at 600 Torr. At each pressure, the branching ratio of C2H5ONO2 formation increases with the decrease of temp. The following parametrization equation has been derived in the pressure and temp. ranges of the study: β(P,T) (%) = (3.88 × 10-3·P (Torr) + 0.365)·(1 + 1500(1/T - 1/298)). The atm. implication of the results obtained is briefly discussed, in particular the impact of β on the evolution of Et nitrate in urban plumes.
- 30Butkovskaya, N. I.; Kukui, A.; Le Bras, G. Pressure Dependence of Iso-Propyl Nitrate Formation in the i-C3H7O2 + NO Reaction. Z. Für Phys. Chem. 2010, 224 (7–8), 1025– 1038, DOI: 10.1524/zpch.2010.6139Google ScholarThere is no corresponding record for this reference.
- 31Atkinson, R.; Aschmann, S. M.; Carter, W. P. L.; Winer, A. M.; Pitts, J. N. Alkyl Nitrate Formation from the Nitrogen Oxide (NOx)-Air Photooxidations of C2-C8 n-Alkanes. J. Phys. Chem. 1982, 86 (23), 4563– 4569, DOI: 10.1021/j100220a022Google Scholar31Alkyl nitrate formation from the nitrogen oxide (NOx)-air photooxidations of C2-C8 n-alkanesAtkinson, Roger; Aschmann, Sara M.; Carter, William P. L.; Winer, Arthur M.; Pitts, James N., Jr.Journal of Physical Chemistry (1982), 86 (23), 4563-9CODEN: JPCHAX; ISSN:0022-3654.The yields of alkyl nitrates formed in the NOx-air photooxidns. of the homologous series of n-alkanes from C2H6 [74-84-0] through n-octane [111-65-9] were detd. at 299 ± 2 K and 735 torr total pressure for different chem. systems. Alkyl peroxy radicals were generated by reaction of the n-alkanes with OH radicals (generated from the photolysis of Me nitrite in air) or Cl atoms (from photolysis of Cl in air). The alkyl nitrate yields obtained from the 2 systems, cor. for secondary reactions, were in agreement within the exptl. errors and increased monotonically with the C no. of the n-alkane, from ≤1% for C2H6 to ∼333% for n-octane, with the yields apparently approaching a limit of ∼35% for larger n-alkanes. The relative yields of the various secondary alkyl nitrate isomers in the n-pentane through n-octane systems were in good agreement with those expected from OH radical or Cl atom reaction with the corresponding secondary C-H bonds. However, the relative yields of the primary alkyl nitrates in the C3H8 and C4H10 systems were lower than expected by a factor of ∼2. The data are consistent with the alkyl nitrates being formed almost entirely from the reaction of peroxy radicals with NO, and the ratios of the cor. alkyl nitrate yields thus reflect the fraction of RO2 radicals which react with NO to form alkyl nitrates. These nitrate yields from the reaction of RO2 radicals with NO are important inputs into chem. computer models of the atm. NOx-air photooxidns. of the larger n-alkanes.
- 32Piletic, I. R.; Edney, E. O.; Bartolotti, L. J. Barrierless Reactions with Loose Transition States Govern the Yields and Lifetimes of Organic Nitrates Derived from Isoprene. J. Phys. Chem. A 2017, 121 (43), 8306– 8321, DOI: 10.1021/acs.jpca.7b08229Google Scholar32Barrierless Reactions with Loose Transition States Govern the Yields and Lifetimes of Organic Nitrates Derived from IsoprenePiletic, Ivan R.; Edney, Edward O.; Bartolotti, Libero J.Journal of Physical Chemistry A (2017), 121 (43), 8306-8321CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The chem. reaction mechanism of NO addn. to two β and δ isoprene hydroxy-peroxy radical isomers is examd. in detail using d. functional theory, coupled cluster methods, and the energy resolved master equation formalism to provide ests. of rate consts. and org. nitrate yields. At the M06-2x/aug-cc-pVTZ level, the potential energy surfaces of NO reacting with β-(1,2)-HO-IsopOO• and δ-Z-(1,4)-HO-IsopOO• possess barrierless reactions that produce alkoxy radicals/NO2 and org. nitrates. The nudged elastic band method was used to discover a loosely bound van der Waals (vdW) complex between NO2 and the alkoxy radical that is present in both exit reaction channels. Semiempirical master equation calcns. show that the β org. nitrate yield is 8.5 ± 3.7%. Addnl., a relatively low barrier to C-C bond scission was discovered in the β-vdW complex that leads to direct HONO formation in the gas phase with a yield of 3.1 ± 1.3%. The δ isomer produces a looser vdW complex with a smaller dissocn. barrier and a larger isomerization barrier, giving a 2.4 ± 0.8% org. nitrate yield that is relatively pressure and temp. insensitive. By considering all of these pathways, the first-generation NOx recycling efficiency from isoprene org. nitrates is estd. to be 21% and is expected to increase with decreasing NOx concn.
- 33Aschmann, S. M.; Arey, J.; Atkinson, R. Atmospheric Chemistry of Selected Hydroxycarbonyls. J. Phys. Chem. A 2000, 104 (17), 3998– 4003, DOI: 10.1021/jp9939874Google Scholar33Atmospheric Chemistry of Selected HydroxycarbonylsAschmann, Sara M.; Arey, Janet; Atkinson, RogerJournal of Physical Chemistry A (2000), 104 (17), 3998-4003CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Using a relative rate method, rate consts. have been measured at 296 ± 2 K for the gas-phase reactions of the OH radical with 1-hydroxy-2-butanone, 3-hydroxy-2-butanone, 1-hydroxy-3-butanone, 1-hydroxy-2-methyl-3-butanone, 3-hydroxy-3-methyl-2-butanone, and 4-hydroxy-3-hexanone, with rate consts. (in units of 10-12 cm3 mol.-1 s-1) of 7.7 ± 1.7, 10.3 ± 2.2, 8.1 ± 1.8, 16.2 ± 3.4, 0.94 ± 0.37, and 15.1 ± 3.1, resp., where the error limits include the estd. overall uncertainty in the rate const. for the ref. compd. Rate consts. were also measured for reactions with NO3 radicals and O3. Rate consts. for the NO3 radical reactions (in units of 10-16 cm3 mol.-1 s-1) were 1-hydroxy-2-butanone, <9; 3-hydroxy-2-butanone, 6.5 ± 2.2; 1-hydroxy-3-butanone, <22; 1-hydroxy-2-methyl-3-butanone, <22; 3-hydroxy-3-methyl-2-butanone, <2; and 4-hydroxy-3-hexanone, 12 ± 4, where the error limits include the estd. overall uncertainties in the rate consts. for the ref. compds. No reactions with O3 were obsd., and upper limits to the rate consts. of <1.1 × 10-19 cm3 mol.-1 s-1 were derived for all six hydroxycarbonyls. The dominant tropospheric loss process for the hydroxycarbonyls studied here is calcd. to be by reaction with the OH radical.
- 34Barry, J. T.; Berg, D. J.; Tyler, D. R. Radical Cage Effects: The Prediction of Radical Cage Pair Recombination Efficiencies Using Microviscosity Across a Range of Solvent Types. J. Am. Chem. Soc. 2017, 139 (41), 14399– 14405, DOI: 10.1021/jacs.7b04499Google Scholar34Radical Cage Effects: The Prediction of Radical Cage Pair Recombination Efficiencies Using Microviscosity Across a Range of Solvent TypesBarry, Justin T.; Berg, Daniel J.; Tyler, David R.Journal of the American Chemical Society (2017), 139 (41), 14399-14405CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)This study reports a method for correlating the radical recombination efficiencies (FcP) of geminate radical cage pairs to the properties of the solvent. Although bulk viscosity (macroviscosity) is typically used to predict or interpret radical recombination efficiencies, the work reported here shows that microviscosity is a much better parameter. The use of microviscosity is valid over a range of different solvent system types, including nonpolar, arom., polar, and hydrogen bonding solvents. In addn., the relationship of FcP to microviscosity holds for solvent systems contg. mixts. of these solvent types. The microviscosities of the solvent systems were straightforwardly detd. by measuring the diffusion coeff. of an appropriate probe by NMR DOSY spectroscopy. By using solvent mixts., selective solvation was shown to not affect the correlation between FcP and microviscosity. In addn., neither solvent polarity nor radical rotation affects the correlation between FcP and the microviscosity.
- 35Pryor, W. A.; Squadrito, G. L. The Chemistry of Peroxynitrite: A Product from the Reaction of Nitric Oxide with Superoxide. Am. J. Physiol.-Lung Cell. Mol. Physiol. 1995, 268 (5), L699– L722, DOI: 10.1152/ajplung.1995.268.5.L699Google ScholarThere is no corresponding record for this reference.
- 36Miyamoto, H.; Yampolski, Y.; Young, C. L. IUPAC-NIST Solubility Data Series. 103. Oxygen and Ozone in Water, Aqueous Solutions, and Organic Liquids (Supplement to Solubility Data Series Volume 7). J. Phys. Chem. Ref. Data 2014, 43 (3), 033102, DOI: 10.1063/1.4883876Google ScholarThere is no corresponding record for this reference.
- 37Sprague, M. K.; Garland, E. R.; Mollner, A. K.; Bloss, C.; Bean, B. D.; Weichman, M. L.; Mertens, L. A.; Okumura, M.; Sander, S. P. Kinetics of n-Butoxy and 2-Pentoxy Isomerization and Detection of Primary Products by Infrared Cavity Ringdown Spectroscopy. J. Phys. Chem. A 2012, 116 (24), 6327– 6340, DOI: 10.1021/jp212136rGoogle Scholar37Kinetics of n-Butoxy and 2-Pentoxy Isomerization and Detection of Primary Products by Infrared Cavity Ringdown SpectroscopySprague, Matthew K.; Garland, Eva R.; Mollner, Andrew K.; Bloss, Claire; Bean, Brian D.; Weichman, Marissa L.; Mertens, Laura A.; Okumura, Mitchio; Sander, Stanley P.Journal of Physical Chemistry A (2012), 116 (24), 6327-6340CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The primary products of n-butoxy and 2-pentoxy isomerization in the presence and absence of O2 have been detected using pulsed laser photolysis-cavity ring down spectroscopy (PLP-CRDS). Alkoxy radicals n-butoxy and 2-pentoxy were generated by photolysis of alkyl nitrite precursors (Bu nitrite or 2-pentyl nitrite, resp.), and the isomerization products with and without O2 were detected by IR cavity ring down spectroscopy 20 μs after the photolysis. We report the mid-IR OH stretch (ν1) absorption spectra for δ-HO-1-C4H8·, δ-HO-1-C4H8OO·, δ-HO-1-C5H10·, and δ-HO-1-C5H10OO·. The obsd. ν1 bands are similar in position and shape to the related alcs. (n-butanol and 2-pentanol), although the HOROO· absorption is slightly stronger than the HOR· absorption. We detd. the rate of isomerization relative to reaction with O2 for the n-butoxy and 2-pentoxy radicals by measuring the relative ν1 absorbance of HOROO· as a function of [O2]. At 295 K and 670 Torr of N2 or N2/O2, we found rate const. ratios of kisom/kO2 = 1.7 (±0.1) × 1019 cm-3 for n-butoxy and kisom/kO2 = 3.4(±0.4) × 1019 cm-3 for 2-pentoxy (2σ uncertainty). Using currently known rate consts. kO2 , we est. isomerization rates of kisom = 2.4 (±1.2) × 105 s-1 and kisom ≈ 3 × 105 s-1 for n-butoxy and 2-pentoxy radicals, resp., where the uncertainties are primarily due to uncertainties in kO2 . Because isomerization is predicted to be in the high pressure limit at 670 Torr, these relative rates are expected to be the same at atm. pressure. Our results include corrections for prompt isomerization of hot nascent alkoxy radicals as well as reaction with background NO and unimol. alkoxy decompn. We est. prompt isomerization yields under our conditions of 4 ± 2% and 5 ± 2% for n-butoxy and 2-pentoxy formed from photolysis of the alkyl nitrites at 351 nm. Our measured relative rate values are in good agreement with and more precise than previous end-product anal. studies conducted on the n-butoxy and 2-pentoxy systems. We show that reactions typically neglected in the anal. of alkoxy relative kinetics (decompn., recombination with NO, and prompt isomerization) may need to be included to obtain accurate values of kisom/kO2 .
- 38Kamath, D.; Mezyk, S. P.; Minakata, D. Elucidating the Elementary Reaction Pathways and Kinetics of Hydroxyl Radical-Induced Acetone Degradation in Aqueous Phase Advanced Oxidation Processes. Environ. Sci. Technol. 2018, 52 (14), 7763– 7774, DOI: 10.1021/acs.est.8b00582Google Scholar38Elucidating the Elementary Reaction Pathways and Kinetics of Hydroxyl Radical-Induced Acetone Degradation in Aqueous Phase Advanced Oxidation ProcessesKamath, Divya; Mezyk, Stephen P.; Minakata, DaisukeEnvironmental Science & Technology (2018), 52 (14), 7763-7774CODEN: ESTHAG; ISSN:0013-936X. (American Chemical Society)Advanced oxidn. processes (AOPs) that produce highly reactive hydroxyl radicals are promising methods to destroy aq. org. contaminants. Hydroxyl radicals react rapidly and nonselectively with org. contaminants and degrade them into intermediates and transformation byproducts. Past studies have indicated that peroxyl radical reactions are responsible for the formation of many intermediate radicals and transformation byproducts. However, complex peroxyl radical reactions that produce identical transformation products make it difficult to exptl. study the elementary reaction pathways and kinetics. In this study, we used ab initio quantum mech. calcns. to identify the thermodynamically preferable elementary reaction pathways of hydroxyl radical-induced acetone degrdn. by calcg. the free energies of the reaction and predicting the corresponding reaction rate consts. by calcg. the free energies of activation. In addn., we solved the ordinary differential equations for each species participating in the elementary reactions to predict the concn. profiles for acetone and its transformation byproducts in an aq. phase UV/hydrogen peroxide AOP. Our ab initio quantum mech. calcns. found an insignificant contribution of Russell reaction mechanisms of peroxyl radicals, but significant involvement of HO2• in the peroxyl radical reactions. The predicted concn. profiles were compared with expts. in the literature, validating our elementary reaction-based kinetic model.
- 39Elford, P. E.; Roberts, B. P. EPR Studies of the Formation and Transformation of Isomeric Radicals [C3H5O]•. Rearrangement of the Allyloxyl Radical in Non-Aqueous Solution Involving a Formal 1,2-Hydrogen-Atom Shift Promoted by Alcohols. J. Chem. Soc., Perkin Trans. 1996, 2 (11), 2247– 2256, DOI: 10.1039/P29960002247Google ScholarThere is no corresponding record for this reference.
- 40Konya, K. G.; Paul, T.; Lin, S.; Lusztyk, J.; Ingold, K. U. Laser Flash Photolysis Studies on the First Superoxide Thermal Source. First Direct Measurements of the Rates of Solvent-Assisted 1,2-Hydrogen Atom Shifts and a Proposed New Mechanism for This Unusual Rearrangement. J. Am. Chem. Soc. 2000, 122 (31), 7518– 7527, DOI: 10.1021/ja993570bGoogle Scholar40Laser Flash Photolysis Studies on the First Superoxide Thermal Source. First Direct Measurements of the Rates of Solvent-Assisted 1,2-Hydrogen Atom Shifts and a Proposed New Mechanism for This Unusual RearrangementKonya, Klara G.; Paul, Thomas; Lin, Shuqiong; Lusztyk, Janusz; Ingold, K. U.Journal of the American Chemical Society (2000), 122 (31), 7518-7527CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The thermal decompn. of bis(4-carboxybenzyl)hyponitrite (I; SOTS-1) in aerated water under physiol. conditions has previously been shown to give the superoxide radical anion in a yield of 40 mol % (Ingold, K. U.; et al. J. Am. Chem. Soc. 1997, 119, 12364). The abs. kinetics of the elementary reactions involved in the cascade of events leading from the first-formed water-sol. benzyloxyl radical to superoxide have been detd. by laser flash photolysis. On the basis of these kinetics it is concluded that SOTS-1 will be suitable for studies of superoxide-induced oxidative stress in most biol. systems. A water-assisted 1,2-H shift converting benzyloxyl into the benzyl ketyl radical is an important step in the above reaction cascade. The kinetics of the 1,2-H shift assisted by H2O, D2O, and a no. of nucleophilic alcs. have been measured for the first time. These data have led to a proposed new mechanism involving the initial formation of a ketyl radical anion and an oxonium cation which generally collapse to give the neutral ketyl radical as the first observable product on the time scale of our expts. (ca. 80 ns).
- 41Schuchmann, H.-P.; Sonntag, C. v. Methylperoxyl Radicals: A Study of the γ-Radiolysis of Methane in Oxygenated Aqueous Solutions. Z. Für Naturforschung B 1984, 39 (2), 217– 221, DOI: 10.1515/znb-1984-0217Google ScholarThere is no corresponding record for this reference.
- 42Schuchmann, H.-P.; von Sonntag, C. Photolysis at 185 nm of Dimethyl Ether in Aqueous Solution: Involvement of the Hydroxymethyl Radical. J. Photochem. 1981, 16 (3), 289– 295, DOI: 10.1016/0047-2670(81)80039-1Google Scholar42Photolysis at 185 nm of dimethyl ether in aqueous solution: involvement of the hydroxymethyl radicalSchuchmann, Heinz Peter; Von Sonntag, ClemensJournal of Photochemistry (1981), 16 (4), 289-95CODEN: JPCMAE; ISSN:0047-2670.The photolysis of Me2O in aq. soln. at 185 nm yielded CH4, H2, MeOH, (MeOCH2)2, HCHO, MeOCH2CH2OH, (HOCH2)2, EtOH, and MeOEt. These products are explained by three primary processes (formation of CH3O• + •CH3; CH3OCH2• + •H; and CH2O + CH4), the rearrangement process (CH3O•→•CH2OH) known to be undergone by alkoxyl radicals in aq. soln. and subsequent free-radical reactions. In aq. solns. the quantum yield of primary processes leading to products is smaller by about an order of magnitude than those in cyclohexane solns. or those previously found with similar ethers as pure liqs. This apparently means that water as a solvent has a quenching effect. In aq. solns. there is an excited species which is reactive towards N2O and a proton, leading to the formation of N2 and H2 resp. Free hydrated electrons generated by photoionization do not appear to be involved in these reactions.
- 43von Sonntag, C.; Schuchmann, H.-P. The Elucidation of Peroxyl Radical Reactions in Aqueous Solution with the Help of Radiation-Chemical Methods. Angew. Chem., Int. Ed. Engl. 1991, 30 (10), 1229– 1253, DOI: 10.1002/anie.199112291Google ScholarThere is no corresponding record for this reference.
- 44Fernández-Ramos, A.; Zgierski, M. Z. Theoretical Study of the Rate Constants and Kinetic Isotope Effects of the 1,2-Hydrogen-Atom Shift of Methoxyl and Benzyloxyl Radicals Assisted by Water. J. Phys. Chem. A 2002, 106 (44), 10578– 10583, DOI: 10.1021/jp020917fGoogle Scholar44Theoretical Study of the Rate Constants and Kinetic Isotope Effects of the 1,2-Hydrogen-Atom Shift of Methoxyl and Benzyloxyl Radicals Assisted by WaterFernandez-Ramos, Antonio; Zgierski, Marek Z.Journal of Physical Chemistry A (2002), 106 (44), 10578-10583CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Rate consts. and kinetic isotope effects for the 1,2-H shift of methoxyl and benzyloxyl radicals were studied in the presence of water mols. The electronic structure calcns. were carried out at the UB3LYP/6-31G* level, and the dynamics calcns. were performed using the variational transition state theory with semiclassical multidimensional corrections for tunneling. The study deals with 1:1 and 1:2 radical-water complexes in the gas phase. It was found that water catalyzes these rearrangement reactions by forming a bridge contg. two water mols. The dynamics calcns. show that the methoxyl-water complexes react only very slowly. The rate consts. for 1:2 complexes with the benzyloxyl radical are in relative good agreement with the results of laser flash photolysis. The kinetic isotope effects calcd. using heavy water indicate that tunneling makes an important contribution in the 1:1 complex, whereas the contribution due to vibrations is more important for the 1:2 complexes. In both cases, the kinetic isotope effects are substantial. It is concluded that the 1,2-H shift for both radicals is catalyzed by two water mols. through a mechanism that involves the formation of a preliminary 1:1 complex.
- 45Sander, R. Compilation of Henry’s Law Constants (Version 4.0) for Water as Solvent. Atmospheric Chem. Phys. 2015, 15 (8), 4399– 4981, DOI: 10.5194/acp-15-4399-2015Google ScholarThere is no corresponding record for this reference.
- 46Mack, J.; Bolton, J. R. Photochemistry of Nitrite and Nitrate in Aqueous Solution: A Review. J. Photochem. Photobiol. Chem. 1999, 128 (1), 1– 13, DOI: 10.1016/S1010-6030(99)00155-0Google Scholar46Photochemistry of nitrite and nitrate in aqueous solution: a reviewMack, John; Bolton, James R.Journal of Photochemistry and Photobiology, A: Chemistry (1999), 128 (1-3), 1-14CODEN: JPPCEJ; ISSN:1010-6030. (Elsevier Science S.A.)It has long been known that the photolysis of nitrite and nitrate solns. results in the formation of OH radicals. The mechanism of NO3- photolysis has been the subject of considerable controversy in the literature, however. This review summarizes the exptl. work on NO2- and NO3- photolysis in the context of recent advances in the understanding of the chem. of the peroxynitrite anion (ONOO-) in biol. expts. ONOO- has been found to play a far more significant role in the overall reaction mechanism of NO3- photolysis than had previously been suspected. Research on NO2- and NO3- photolysis, as a pathway to the destruction of org. contaminants in natural waters, is summarized. The possible impact of NO2- and NO3- on Advanced Oxidn. Technologies (AOTs), in which OH radicals are used to initiate the destruction of hazardous org. pollutants in drinking water and industrial waste streams, is explored. A review with 133 refs.
- 47Méreau, R.; Rayez, M.-T.; Caralp, F.; Rayez, J.-C. Isomerisation Reactions of Alkoxy Radicals: Theoretical Study and Structure–Activity Relationships. Phys. Chem. Chem. Phys. 2003, 5 (21), 4828– 4833, DOI: 10.1039/B307708JGoogle 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 1 publications.
- Federica Valentini, Ivo Allegrini, Irene Colasanti, Camilla Zaratti, Andrea Macchia, Cristiana Barandoni, Anna Neri. Preliminary Assessment of Air Pollution in the Archaeological Museum of Naples (Italy): Long Term Monitoring of Nitrogen Dioxide and Nitrous Acid. Air 2025, 3
(2)
, 12. https://doi.org/10.3390/air3020012
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. : Time-resolved mass spectrometric signals for n-pentyl nitrite photolysis products, in three phases: (a) the gas phase, (b) the condensed organic phase, and (3) the aqueous phase. All measurements use the NH4+ CIMS, and formulas listed are those of the analytes (products are detected as the complexes with NH4+). Signals are smoothed over 15-s intervals and normalized to the dilution tracer, acetonitrile; t = 0 refers to the time at which UV lights are switched on.
Figure 2
Figure 2. : Simplified reaction mechanism for the photolysis of n-pentyl nitrite. Detected closed-shell products are highlighted, with colors corresponding to the products shown in Figure 1.
References
This article references 47 other publications.
- 1Hoffmann, 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 (42), 11776– 11781, DOI: 10.1073/pnas.16063201131An 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.
- 2George, I. J.; Abbatt, J. P. D. Heterogeneous Oxidation of Atmospheric Aerosol Particles by Gas-Phase Radicals. Nat. Chem. 2010, 2 (9), 713– 722, DOI: 10.1038/nchem.8062Heterogeneous oxidation of atmospheric aerosol particles by gas-phase radicalsGeorge, I. J.; Abbatt, J. P. D.Nature Chemistry (2010), 2 (9), 713-722CODEN: NCAHBB; ISSN:1755-4330. (Nature Publishing Group)A review concerning the heterogeneous oxidn. of org. and inorg. constituents of atm. aerosols by gas-phase radicals is given, focusing on the kinetics and reaction mechanisms and how they affect particle physicochem. properties (compn., size, d., hygroscopicity). Potential atm. impacts include: release of chem. reactive gases, e.g., halogens, aldehydes and org. acids; reactive loss of particle-borne mol. tracer and toxic species; and enhanced hygroscopic properties of aerosols which may improve their ability to form cloud droplets. Topics discussed include: radical uptake on atm. relevant surfaces; oxidn. of condensed-phase org. and inorg. compds.; particle physicochem. property modification; atm. implications; and outlook.
- 3Lambe, A. T.; Cappa, C. D.; Massoli, P.; Onasch, T. B.; Forestieri, S. D.; Martin, A. T.; Cummings, M. J.; Croasdale, D. R.; Brune, W. H.; Worsnop, D. R.; Davidovits, P. Relationship between Oxidation Level and Optical Properties of Secondary Organic Aerosol. Environ. Sci. Technol. 2013, 47 (12), 6349– 6357, DOI: 10.1021/es401043j3Relationship between Oxidation Level and Optical Properties of Secondary Organic AerosolLambe, Andrew T.; Cappa, Christopher D.; Massoli, Paola; Onasch, Timothy B.; Forestieri, Sara D.; Martin, Alexander T.; Cummings, Molly J.; Croasdale, David R.; Brune, William H.; Worsnop, Douglas R.; Davidovits, PaulEnvironmental Science & Technology (2013), 47 (12), 6349-6357CODEN: ESTHAG; ISSN:0013-936X. (American Chemical Society)Brown carbon (BrC), which may include secondary org. aerosol (SOA), can be a significant climate-forcing agent through its optical absorption properties; however, the overall contribution of SOA to BrC remains poorly understood. This work examd. correlations between oxidn. level and SOA optical properties. SOA was generated in a flow reactor in the absence of NOx by OH- oxidn. of gaseous precursors as surrogates for anthropogenic (naphthalene, tricyclo[5.2.1.02,6]decane), biomass burning (guaiacol), and biogenic (α-pinene) emissions. SOA chem. compn. was characterized by time-of-flight aerosol mass spectrometry. SOA mass-specific absorption cross sections (MAC) and refractive indexes were calcd. from real-time cavity ring-down photo-acoustic spectrometry measurements at 405 and 532 nm and from UV-vis spectrometry measurements of methanol exts. of filter-collected particles (300-600 nm). At 405 nm, SOA MAC values and imaginary refractive indexes increased with increasing oxidn. level and decreased with increasing wavelength, leading to negligible absorption at 532 nm. SOA real refractive indexes decreased with increasing oxidn. level. A literature study comparison suggested that under typical polluted conditions, the effect of NOx on SOA absorption was small. SOA may significantly contribute to atm. BrC; the magnitude depends on precursor type and oxidn. level.
- 4Galeazzo, T.; Valorso, R.; Li, Y.; Camredon, M.; Aumont, B.; Shiraiwa, M. Estimation of Secondary Organic Aerosol Viscosity from Explicit Modeling of Gas-Phase Oxidation of Isoprene and α-Pinene. Atmospheric Chem. Phys. 2021, 21 (13), 10199– 10213, DOI: 10.5194/acp-21-10199-20214Estimation of secondary organic aerosol viscosity from explicit modeling of gas-phase oxidation of isoprene and α-pineneGaleazzo, Tommaso; Valorso, Richard; Li, Ying; Camredon, Marie; Aumont, Bernard; Shiraiwa, ManabuAtmospheric Chemistry and Physics (2021), 21 (13), 10199-10213CODEN: ACPTCE; ISSN:1680-7324. (Copernicus Publications)Secondary org. aerosols (SOA) are major components of atm. fine particulate matter, affecting climate and air quality. Mounting evidence exists that SOA can adopt glassy and viscous semisolid states, impacting formation and partitioning of SOA. In this study, we apply the GECKO-A (Generator of Explicit Chem. and Kinetics of Orgs. in the Atm.) model to conduct explicit chem. modeling of isoprene photooxidn. and α-pinene ozonolysis and their subsequent SOA formation. The detailed gas-phase chem. schemes from GECKO-A are implemented into a box model and coupled to our recently developed glass transition temp. parameterizations, allowing us to predict SOA viscosity. The effects of chem. compn., relative humidity, mass loadings and mass accommodation on particle viscosity are investigated in comparison with measurements of SOA viscosity. The simulated viscosity of isoprene SOA agrees well with viscosity measurements as a function of relative humidity, while the model underestimates viscosity of α-pinene SOA by a few orders of magnitude. This difference may be due to missing processes in the model, including autoxidn. and particle-phase reactions, leading to the formation of high-molar-mass compds. that would increase particle viscosity. Addnl. simulations imply that kinetic limitations of bulk diffusion and redn. in mass accommodation coeff. may play a role in enhancing particle viscosity by suppressing condensation of semi-volatile compds. The developed model is a useful tool for anal. and investigation of the interplay among gas-phase reactions, particle chem. compn. and SOA phase state.
- 5Kroll, J. H.; Seinfeld, J. H. Chemistry of Secondary Organic Aerosol: Formation and Evolution of Low-Volatility Organics in the Atmosphere. Atmos. Environ. 2008, 42 (16), 3593– 3624, DOI: 10.1016/j.atmosenv.2008.01.0035Chemistry of secondary organic aerosol: Formation and evolution of low-volatility organics in the atmosphereKroll, Jesse H.; Seinfeld, John H.Atmospheric Environment (2008), 42 (16), 3593-3624CODEN: AENVEQ; ISSN:1352-2310. (Elsevier Ltd.)A review is given. Secondary org. aerosol (SOA), particulate matter composed of compds. formed from the atm. transformation of org. species, accounts for a substantial fraction of tropospheric aerosol. The formation of low-volatility (semivolatile and possibly nonvolatile) compds. that make up SOA is governed by a complex series of reactions of a large no. of org. species, so the exptl. characterization and theor. description of SOA formation presents a substantial challenge. We outline what is known about the chem. of formation and continuing transformation of low-volatility species in the atm. The primary focus is chem. processes that can change the volatility of org. compds.: (1) oxidn. reactions in the gas phase, (2) reactions in the particle phase, and (3) continuing chem. (in either phase) over several generations. Gas-phase oxidn. reactions can reduce volatility by the addn. of polar functional groups or increase it by the cleavage of carbon-carbon bonds; key branch points that control volatility are the initial attack of the oxidant, reactions of alkylperoxy (RO2) radicals, and reactions of alkoxy (RO) radicals. Reactions in the particle phase include oxidn. reactions as well as accretion reactions, non-oxidative processes leading to the formation of high-mol.-wt. species. Org. C in the atm. is continually subject to reactions in the gas and particle phases throughout its atm. lifetime (until lost by phys. deposition or oxidized to CO or CO2), implying continual changes in volatility over the timescales of several days. The volatility changes arising from these chem. reactions must be parameterized and included in models in order to gain a quant. and predictive understanding of SOA formation.
- 6Calvert, J. G.; Derwent, R. G.; Orlando, J. J.; Wallington, T. J.; Tyndall, G. S. Mechanisms of Atmospheric Oxidation of the Alkanes; Oxford University Press: Oxford, NY, 2008.There is no corresponding record for this reference.
- 7Atkinson, R.; Arey, J. Atmospheric Degradation of Volatile Organic Compounds. Chem. Rev. 2003, 103 (12), 4605– 4638, DOI: 10.1021/cr02064207Atmospheric Degradation of Volatile Organic CompoundsAtkinson, Roger; Arey, JanetChemical Reviews (Washington, DC, United States) (2003), 103 (12), 4605-4638CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review concerning the removal of biogenic and anthropogenic volatile org. compds. (VOC) emitted to the atm. through phys. (deposition) or chem. processes, or their atm. transformation is given. Topics discussed include: tropospheric VOC transformation processes (initial reactions and lifetimes, reaction mechanisms); atm. chem. of alkanes (kinetic data for initial OH- and NO3- reactions, reaction mechanism); atm. chem. of alkenes (rate consts. for initial reaction of alkenes with OH-, NO3-, and O3; mechanism of the OH- reaction; NO3- reaction; reaction with O3); arom. hydrocarbons (kinetics of OH- reactions, reactions of phenols, reactions of unsatd. 1,4-dicarbonyls and di-unsatd. 1,6-dicarbonyls); atm. reactions of oxygenated VOC (aldehydes, ketones, aliph. alcs., ethers, alkyl nitrates); and conclusions and future research needs (exptl. studies, theor. studies and crit. reviews and evaluations).
- 8Calvert, J. G.; Mellouki, A.; Orlando, J. J. Mechanisms of Atmospheric Oxidation of the Oxygenates; Oxford University Press: Oxford, NY, 2011. DOI: 10.1093/oso/9780195365818.005.0001 .There is no corresponding record for this reference.
- 9Leviss, D. H.; Van Ry, D. A.; Hinrichs, R. Z. Multiphase Ozonolysis of Aqueous α-Terpineol. Environ. Sci. Technol. 2016, 50 (21), 11698– 11705, DOI: 10.1021/acs.est.6b036129Multiphase Ozonolysis of Aqueous α-TerpineolLeviss, Dani H.; Van Ry, Daryl A.; Hinrichs, Ryan Z.Environmental Science & Technology (2016), 50 (21), 11698-11705CODEN: ESTHAG; ISSN:0013-936X. (American Chemical Society)Multiphase ozonolysis of aq. orgs. presents a potential pathway for the formation of aq. secondary org. aerosol (aqSOA). We investigated the multiphase ozonolysis of α-terpineol, an oxygenated deriv. of limonene, and found that the reaction products and kinetics differ from the gas-phase ozonolysis of α-terpineol. One- and two-dimensional NMR spectroscopies along with GC-MS identified the aq. ozonolysis reaction products as trans- and cis-lactols [4-(5-hydroxy-2,2-dimethyltetrahydrofuran-3-yl)butan-2-one] and a lactone [4-hydroxy-4-methyl-3-(3-oxobutyl)-valeric acid gamma-lactone], which accounted for 46%, 27%, and 20% of the obsd. products, resp. Hydrogen peroxide was also formed in 10% yield consistent with a mechanism involving decompn. of hydroxyl hydroperoxide intermediates followed by hemiacetal ring closure. Multiphase reaction kinetics at gaseous ozone concns. of 131, 480, and 965 parts-per-billion were analyzed using a resistance model of net ozone uptake and found the second-order rate coeff. for the aq. reaction of α-terpineol + O3 to be 9.9(±3.3) × 106 M-1 s-1. Multiphase ozonolysis will therefore be competitive with multiphase oxidn. by hydroxyl radicals (OH) and ozonolysis of gaseous α-terpineol. We also measured product yields for the heterogeneous ozonolysis of α-terpineol adsorbed on glass, NaCl, and kaolinite, and identified the same three major products but with an increasing lactone yield of 33, 49, and 55% on these substrates, resp.
- 10Akimoto, H.; Hirokawa, J. Atmospheric Multiphase Chemistry: Fundamentals of Secondary Aerosol Formation; John Wiley & Sons, 2020. DOI: 10.1002/9781119422419 .There is no corresponding record for this reference.
- 11Thornberry, T.; Abbatt, J. P. D. Heterogeneous Reaction of Ozone with Liquid Unsaturated Fatty Acids: Detailed Kinetics and Gas-Phase Product Studies. Phys. Chem. Chem. Phys. 2004, 6 (1), 84– 93, DOI: 10.1039/b310149e11Heterogeneous reaction of ozone with liquid unsaturated fatty acids: detailed kinetics and gas-phase product studiesThornberry, T.; Abbatt, J. P. D.Physical Chemistry Chemical Physics (2004), 6 (1), 84-93CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)Detailed kinetic and product yield studies have been performed for the heterogeneous reaction between gas-phase ozone and three liq. fatty acids using a coated-wall flow tube and chem. ionization mass spectrometry. Gas-surface reaction probabilities for ozone loss of (8.0 ±1.0) × 10-4, (1.3 ±0.1) × 10-3, and (1.8 ±0.2) × 10-3 have been measured at room temp. (298 K) for oleic acid, linoleic acid, and linolenic acid, resp. The temp. dependence of the uptake coeffs. was found to be small and pos. Comparison of these results to the kinetics of the equiv. gas-phase reactions implies that there is a definite enhancement in the rate for the heterogeneous process due to entropic factors, i.e., due to collisional trapping of ozone in the surface layers of the liq., and a possible effect on the activation energy of the reaction. For linoleic acid, the reaction probability was found to be independent of relative humidity (up to 55%), to ±10%, at 263 K. Volatile reaction products were obsd. using proton-transfer-reaction mass spectrometry. Nonanal was obsd. with a 0.50 (±0.10) yield for the reaction with oleic acid, whereas hexanal and nonenal were obsd. for linoleic acid with 0.25 (±0.05) and 0.29 (±0.05) yields, resp. These results indicate that the primary ozonide formed initially in the reaction can decomp. via two equal probability pathways and that a secondary ozonide is not formed in high yield in the aldehydic channel. These reactions represent a source of oxygenates to the atm. and will modify the hygroscopic properties of aerosols.
- 12Adams, G. E.; Willson, R. L. Pulse Radiolysis Studies on the Oxidation of Organic Radicals in Aqueous Solution. Trans. Faraday Soc. 1969, 65, 2981, DOI: 10.1039/tf969650298112Pulse radiolysis studies on the oxidation of organic radicals in aqueous solutionAdams, Gerald Edward; Willson, Robin LindhopeTransactions of the Faraday Society (1969), 65 (11), 2981-7CODEN: TFSOA4; ISSN:0014-7672.Pulse radiolysis has been used to measure directly the abs. rates of oxidn. by ferricyanide ion of various radicals produced by OH attack on org. solutes. These include mono-, di-, and polyhydroxylic compds., hydroxy acids, polyethylene oxides of mol. wt. 200, 6000, and 20,000 and the amino acid serine. Radicals produced by H abstraction from α-C atoms in alcs. are oxidized at, or near, diffusion-controlled rates, whereas the reactions are much slower for radicals formed by OH-attack elsewhere. The technique has been used to measure the percentage of OH attack at the α-position for a series of straight and branched-chain alcs. O competes with ferricyanide for radical oxidn. The data for O-contg. solns. fit a simple radical-competition scheme which has been used to measure rates of peroxy-radical formation. These approach diffusion-controlled limits.
- 13Schuchmann, M. N.; Von Sonntag, C. The Rapid Hydration of the Acetyl Radical. A Pulse Radiolysis Study of Acetaldehyde in Aqueous Solution. J. Am. Chem. Soc. 1988, 110 (17), 5698– 5701, DOI: 10.1021/ja00225a01913The rapid hydration of the acetyl radical. A pulse radiolysis study of acetaldehyde in aqueous solution.Schuchmann, Man Nien; Von Sonntag, ClemensJournal of the American Chemical Society (1988), 110 (17), 5698-701CODEN: JACSAT; ISSN:0002-7863.In aq. soln. acetaldehyde and its hydrate are in a 0.8:1 equil. Hydroxyl radicals, generated by the pulse radiolysis of N2O-satd. water, react with acetaldehyde about 3 times faster than with the hydrate. The predominant radicals formed are the acetyl radical and its hydrated form, H-abstraction at Me occurring to only about 5-10%. The acetyl radical rapidly hydrates. Its rate of hydration is 2 × 106 times faster than that of acetaldehyde. The hydrated acetyl radical has been monitored by its rapid redn. of tetranitromethane, the formation of O2- in the presence of Oxygen, and its deprotonation at pH 11. The acetyl radical does not oxidize N,N,N',N'-tetramethyl-p-phenylenediamine (TMPD) on the pulse radiolysis time scale. The acetylperoxyl radical (formed in the presence of oxygen) reacts rapidly with TMPD, ascorbate, and O2-; it is the most strongly oxidizing peroxyl radical known so far. Data on the rate of water loss from the hydrated acetyl radical are considerably less accurate, but the rate const. in estd.
- 14Mezyk, S. P.; Elliot, A. J. Pulse Radiolysis of Iodate in Aqueous Solution. J. Chem. Soc. Faraday Trans. 1994, 90 (6), 831– 836, DOI: 10.1039/ft9949000831There is no corresponding record for this reference.
- 15Lee, A. K. Y.; Zhao, R.; Gao, S. S.; Abbatt, J. P. D. Aqueous-Phase OH Oxidation of Glyoxal: Application of a Novel Analytical Approach Employing Aerosol Mass Spectrometry and Complementary Off-Line Techniques. J. Phys. Chem. A 2011, 115 (38), 10517– 10526, DOI: 10.1021/jp204099g15Aqueous-phase OH oxidation of glyoxal: application of a novel analytical approach employing aerosol mass spectrometry and complementary off-line techniquesLee, Alex K. Y.; Zhao, R.; Gao, S. S.; Abbatt, J. P. D.Journal of Physical Chemistry A (2011), 115 (38), 10517-10526CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Aq.-phase chem. of glyoxal may play an important role in the formation of highly oxidized secondary org. aerosol (SOA) in the atm. In this work, we use a novel design of photochem. reactor that allows for simultaneous photo-oxidn. and atomization of a bulk soln. to study the aq.-phase OH oxidn. of glyoxal. By employing both online aerosol mass spectrometry (AMS) and offline ion chromatog. (IC) measurements, glyoxal and some major products including formic acid, glyoxylic acid, and oxalic acid in the reacting soln. were simultaneously quantified. This is the first attempt to use AMS in kinetics studies of this type. The results illustrate the formation of highly oxidized products that likely coexist with traditional SOA materials, thus, potentially improving model predictions of org. aerosol mass loading and degree of oxidn. Formic acid is the major volatile species identified, but the atm. relevance of its formation chem. needs to be further investigated. While successfully quantifying low mol. wt. org. oxygenates and tentatively identifying a reaction product formed directly from glyoxal and hydrogen peroxide, comparison of the results to the offline total org. carbon (TOC) anal. clearly shows that the AMS is not able to quant. monitor all dissolved orgs. in the bulk soln. This is likely due to their high volatility or low stability in the evapd. soln. droplets. This exptl. approach simulates atm. aq. phase processing by conducting oxidn. in the bulk phase, followed by evapn. of water and volatile orgs. to form SOA.
- 16Carrasquillo, A. J.; Hunter, J. F.; Daumit, K. E.; Kroll, J. H. Secondary Organic Aerosol Formation via the Isolation of Individual Reactive Intermediates: Role of Alkoxy Radical Structure. J. Phys. Chem. A 2014, 118 (38), 8807– 8816, DOI: 10.1021/jp506562r16Secondary Organic Aerosol Formation via the Isolation of Individual Reactive Intermediates: Role of Alkoxy Radical StructureCarrasquillo, Anthony J.; Hunter, James F.; Daumit, Kelly E.; Kroll, Jesse H.Journal of Physical Chemistry A (2014), 118 (38), 8807-8816CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The study of the chem. underlying secondary org. aerosol (SOA) formation is complicated by the large no. of reaction pathways and oxidn. generations available to a given precursor species. Here we simplify such complexity to that of a single alkoxy radical (RO), by forming SOA via the direct photolysis of alkyl nitrite (RONO) isomers. Chamber expts. were conducted with 11 C10 RONO isomers to det. how the position of the radical center and branching of the carbon skeleton influences SOA formation. SOA yields served as a probe of RO reactivity, with lower yields indicating that fragmentation reactions dominate and higher yields suggesting the predominance of RO isomerization. The largest yields were from straight-chain isomers, particularly those with radical centers located toward the terminus of the mol. Trends in SOA yields can be explained in terms of two major effects: (1) the relative importance of isomerization and fragmentation reactions, which control the distribution of products, and (2) differences in volatility among the various isomeric products formed. Yields from branched isomers, which were low but variable, provide insight into the degree of fragmentation of the alkoxy radicals; in the case of the two β-substituted alkoxy radicals, fragmentation appears to occur to a greater extent than predicted by structure-activity relationships. Our results highlight how subtle differences in alkoxy radical structure can have major impacts on product yields and SOA formation.
- 17Kessler, S. H.; Nah, T.; Carrasquillo, A. J.; Jayne, J. T.; Worsnop, D. R.; Wilson, K. R.; Kroll, J. H. Formation of Secondary Organic Aerosol from the Direct Photolytic Generation of Organic Radicals. J. Phys. Chem. Lett. 2011, 2 (11), 1295– 1300, DOI: 10.1021/jz200432n17Formation of Secondary Organic Aerosol from the Direct Photolytic Generation of Organic RadicalsKessler, Sean H.; Nah, Theodora; Carrasquillo, Anthony J.; Jayne, John T.; Worsnop, Douglas R.; Wilson, Kevin R.; Kroll, Jesse H.Journal of Physical Chemistry Letters (2011), 2 (11), 1295-1300CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)The immense complexity inherent in secondary org. aerosol (SOA) formation, primarily due to the large no. of oxidn. steps and reaction pathways involved, has limited a detailed understanding of its underlying chem. To simplify such complexity, this work demonstrates SOA formation via photolysis of gas-phase alkyl iodides, generating org. peroxy radicals with known structures. In contrast to std. OH--initiated oxidn. expts., photolytically-initiated oxidn. forms a limited no. of products in a single reactive step. Typical for SOA, yields of aerosol generated from alkyl iodide photolysis depends on aerosol load, indicating the semi-volatile nature of the particulate species. However, aerosols were obsd. to have higher volatility and be less oxidized than previous multi-generational studies of alkane oxidn., suggesting addnl. oxidative steps are necessary to produce oxidized semi-volatile matter in the atm. Despite the relative simplicity of this chem. system, SOA mass spectra are still quite complex, underscoring the wide range of products present in SOA.
- 18Iglesias, E.; Casado, J. Mechanisms of Hydrolysis and Nitrosation Reactions of Alkyl Nitrites in Various Media. Int. Rev. Phys. Chem. 2002, 21 (1), 37– 74, DOI: 10.1080/01442350110092693There is no corresponding record for this reference.
- 19Hunter, J. F.; Carrasquillo, A. J.; Daumit, K. E.; Kroll, J. H. Secondary Organic Aerosol Formation from Acyclic, Monocyclic, and Polycyclic Alkanes. Environ. Sci. Technol. 2014, 48 (17), 10227– 10234, DOI: 10.1021/es502674s19Secondary organic aerosol formation from acyclic, monocyclic, and polycyclic alkanesHunter, James F.; Carrasquillo, Anthony J.; Daumit, Kelly E.; Kroll, Jesse H.Environmental Science & Technology (2014), 48 (17), 10227-10234CODEN: ESTHAG; ISSN:0013-936X. (American Chemical Society)A large no. of org. species emitted into the atm. contain cycloalkyl groups. While cyclic species are believed to be important secondary org. aerosol (SOA) precursors, the specific role of cyclic moieties (particularly for species with multiple or fused rings) remains uncertain. Here we examine the yields and compn. of SOA formed from the reaction of OH with a series of C10 (cyclo)alkanes, with 0-3 rings, in order to better understand the role of multiple cyclic moieties on aerosol formation pathways. A chamber oxidn. technique using high, sustained OH radical concns. was used to simulate long reaction times in the atm. This aging technique leads to higher yields than in previously reported chamber expts. Yields were highest for cyclic and polycyclic precursors, though yield exhibited little dependence on no. of rings. However, the oxygen-to-carbon ratio of the SOA was highest for the polycyclic precursors. These trends are consistent with aerosol formation requiring two generations of oxidn. and 3-4 oxygen-contg. functional groups in order to condense. Cyclic, unbranched structures are protected from fragmentation during the first oxidn. step, with C-C bond scission instead leading to ring opening, efficient functionalization, and high SOA yields. Fragmentation may occur during subsequent oxidn. steps, limiting yields by forming volatile products. Polycyclic structures can undergo multiple ring opening reactions, but do not have markedly higher yields, likely due to enhanced fragmentation in the second oxidn. step. By contrast, C-C bond scission for the linear and branched structures leads to fragmentation prior to condensation, resulting in low SOA yields. The results highlight the key roles of multigenerational chem. and susceptibility to fragmentation in the formation and evolution of SOA.
- 20Zaytsev, A.; Breitenlechner, M.; Koss, A. R.; Lim, C. Y.; Rowe, J. C.; Kroll, J. H.; Keutsch, F. N. Using Collision-Induced Dissociation to Constrain Sensitivity of Ammonia Chemical Ionization Mass Spectrometry (NH4+ CIMS) to Oxygenated Volatile Organic Compounds. Atmospheric Meas. Technol. 2019, 12 (3), 1861– 1870, DOI: 10.5194/amt-12-1861-2019There is no corresponding record for this reference.
- 21Orlando, J. J.; Tyndall, G. S.; Wallington, T. J. The Atmospheric Chemistry of Alkoxy Radicals. Chem. Rev. 2003, 103 (12), 4657– 4690, DOI: 10.1021/cr020527p21The Atmospheric Chemistry of Alkoxy RadicalsOrlando, John J.; Tyndall, Geoffrey S.; Wallington, Timothy J.Chemical Reviews (Washington, DC, United States) (2003), 103 (12), 4657-4689CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review concerning the atm. chem. of alkoxy radicals and rates and mechanisms of various reaction pathways is given. Topics discussed include: exptl. methods used to study alkoxy radical chem.; alkoxy radical chem. (reaction with O2 [CH3O• + O2; C2H5O• + O2; 1-C3H7O•, 2-C2H7O•, 1-C4H9O•, 2-C2H9O•, 3-C5H11O• + O2; CH2ClO•, CFCl2CH2O• + O2; comparison with previous recommendations], dissocn. reactions [unsubstituted alkoxy radicals, β-hydroxyalkoxy radicals, other O-substituted alkoxy radicals, halogenated alkoxy radicals], isomerization reactions, other intramol. reactions [HCl elimination reactions, the ester rearrangement reaction], chem. activation); and conclusions and suggestions for further study.
- 22Atkinson, R. Rate Constants for the Atmospheric Reactions of Alkoxy Radicals: An Updated Estimation Method. Atmos. Environ. 2007, 41 (38), 8468– 8485, DOI: 10.1016/j.atmosenv.2007.07.00222Rate constants for the atmospheric reactions of alkoxy radicals: An updated estimation methodAtkinson, RogerAtmospheric Environment (2007), 41 (38), 8468-8485CODEN: AENVEQ; ISSN:1352-2310. (Elsevier Ltd.)Alkoxy radicals are key intermediates in the atm. degrdns. of volatile org. compds., and can typically undergo reaction with O2, unimol. decompn. or unimol. isomerization. Previous structure-reactivity relationships for the estn. of rate consts. for these processes for alkoxy radicals have been updated to incorporate recent kinetic data from abs. and relative rate studies. Temp.-dependent rate expressions are derived allowing rate consts. for all three of these alkoxy radical reaction pathways to be calcd. at atmospherically relevant temps.
- 23Jenkin, M. E.; Valorso, R.; Aumont, B.; Rickard, A. R. Estimation of Rate Coefficients and Branching Ratios for Reactions of Organic Peroxy Radicals for Use in Automated Mechanism Construction. Atmospheric Chem. Phys. 2019, 19 (11), 7691– 7717, DOI: 10.5194/acp-19-7691-201923Estimation of rate coefficients and branching ratios for reactions of organic peroxy radicals for use in automated mechanism constructionJenkin, Michael E.; Valorso, Richard; Aumont, Bernard; Rickard, Andrew R.Atmospheric Chemistry and Physics (2019), 19 (11), 7691-7717CODEN: ACPTCE; ISSN:1680-7324. (Copernicus Publications)Org. peroxy radicals (RO2), formed from the degrdn. of hydrocarbons and other volatile org. compds. (VOCs), play a key role in tropospheric oxidn. mechanisms. Several competing reactions may be available for a given RO2 radical, the relative rates of which depend on both the structure of RO2 and the ambient conditions. Published kinetics and branching ratio data are reviewed for the bimol. reactions of RO2 with NO, NO2, NO3, OH and HO2; and for their self-reactions and cross-reactions with other RO2 radicals. This information is used to define generic rate coeffs. and structure-activity relationship (SAR) methods that can be applied to the bimol. reactions of a series of important classes of hydrocarbon and oxygenated RO2 radicals. Information for selected unimol. isomerization reactions (i.e. H-atom shift and ring-closure reactions) is also summarized and discussed. The methods presented here are intended to guide the representation of RO2 radical chem. in the next generation of explicit detailed chem. mechanisms.
- 24Dahl, E. E.; Saltzman, E. S.; De Bruyn, W. J. The Aqueous Phase Yield of Alkyl Nitrates from ROO + NO: Implications for Photochemical Production in Seawater. Geophys. Res. Lett. 2003, 30 (6), 1271, DOI: 10.1029/2002GL016811There is no corresponding record for this reference.
- 25Butkovskaya, N.; Kukui, A.; Le Bras, G. Pressure and Temperature Dependence of Methyl Nitrate Formation in the CH3O2 + NO Reaction. J. Phys. Chem. A 2012, 116 (24), 5972– 5980, DOI: 10.1021/jp210710d25Pressure and Temperature Dependence of Methyl Nitrate Formation in the CH3O2 + NO ReactionButkovskaya, Nadezhda; Kukui, Alexandre; Le Bras, GeorgesJournal of Physical Chemistry A (2012), 116 (24), 5972-5980CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The branching ratio β = k1b/k1a for the formation of Me nitrate, CH3ONO2, in the gas-phase CH3O2 + NO reaction, CH3O2 + NO → CH3O + NO2 (1a), CH3O2 + NO → CH3ONO2 (1b), has been detd. over the pressure and temp. ranges 50-500 Torr and 223-300 K, resp., using a turbulent flow reactor coupled with a chem. ionization mass spectrometer. At 298 K, the CH3ONO2 yield has been found to increase linearly with pressure from 0.33 ± 0.16% at 50 Torr to 0.80 ± 0.54% at 500 Torr (errors are 2σ). Decrease of temp. from 300 to 220 K leads to an increase of β by a factor of about 3 in the 100-200 Torr range. These data correspond to a value of β ≈ 1.0 ± 0.7% over the pressure and temp. ranges of the whole troposphere. Atm. concns. of CH3ONO2 roughly estd. using results of this work are in reasonable agreement with those obsd. in polluted environments and significantly higher compared with measurements in upper troposphere and lower stratosphere.
- 26Richards-Henderson, N. K.; Goldstein, A. H.; Wilson, K. R. Large Enhancement in the Heterogeneous Oxidation Rate of Organic Aerosols by Hydroxyl Radicals in the Presence of Nitric Oxide. J. Phys. Chem. Lett. 2015, 6 (22), 4451– 4455, DOI: 10.1021/acs.jpclett.5b0212126Large Enhancement in the Heterogeneous Oxidation Rate of Organic Aerosols by Hydroxyl Radicals in the Presence of Nitric OxideRichards-Henderson, Nicole K.; Goldstein, Allen H.; Wilson, Kevin R.Journal of Physical Chemistry Letters (2015), 6 (22), 4451-4455CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)In the troposphere, the heterogeneous lifetime of an org. mol. in an aerosol exposed to OH- is thought to be weeks, which is orders of magnitude slower than the analogous gas phase reactions (hours). This paper reports an unexpectedly large acceleration in the effective heterogeneous OH- reaction rate in the presence of NO. This 10-50 fold acceleration originates from free radical chain reactions, propagated by alkoxy radicals which form inside the aerosols by the NO reaction with peroxy radicals, which do not appear to produce chain-terminating products (e.g., alkyl nitrates), unlike gas phase mechanisms. A kinetic model, constrained by exptl. data, suggests that in polluted regions, heterogeneous oxidn. plays a much more prominent role in the daily chem. evolution of org. aerosols than previously believed.
- 27Renbaum, L. H.; Smith, G. D. Organic Nitrate Formation in the Radical-Initiated Oxidation of Model Aerosol Particles in the Presence of NOx. Phys. Chem. Chem. Phys. 2009, 11 (36), 8040– 8047, DOI: 10.1039/b909239kThere is no corresponding record for this reference.
- 28Goldstein, S.; Lind, J.; Merenyi, G. Reaction of Organic Peroxyl Radicals with •NO2 and •NO in Aqueous Solution: Intermediacy of Organic Peroxynitrate and Peroxynitrite Species. J. Phys. Chem. A 2004, 108 (10), 1719– 1725, DOI: 10.1021/jp037431z28Reaction of Organic Peroxyl Radicals with •NO2 and •NO in Aqueous Solution: Intermediacy of Organic Peroxynitrate and Peroxynitrite SpeciesGoldstein, Sara; Lind, Johan; Merenyi, GaborJournal of Physical Chemistry A (2004), 108 (10), 1719-1725CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)In this work, we studied the reactions of alkyl peroxyl radicals with •NO2 and •NO using the pulse radiolysis technique. The rate consts. for the reaction of •NO2 with (CH3)2C(OH)CH2OO•, CH3OO•, and c-C5H9OO• vary between 7 × 108 and 1.5 × 109 M-1 s-1. The reaction produces relatively long-lived alkyl peroxynitrates, which are in equil. with the parent radicals and have no appreciable absorption above 270 nm. It is also shown that •NO adds rapidly to (CH3)2C(OH)CH2OO• and CH3OO• to form alkyl peroxynitrites. The rate consts. for these reactions were detd. to be 2.8 × 109 and 3.5 × 109 M-1 s-1, resp. However, in contrast to alkyl peroxynitrates, alkyl peroxynitrites do not accumulate. Rather, they decomp. rapidly via homolysis along the relatively weak O-O bond, initially forming a geminate pair. Most of this pair collapses in the cage to form an alkyl nitrate, RONO2, and about 14% diffuses out as free alkoxyl and •NO2 radicals. A thermokinetic anal. predicts the half-life of CH3OONO in water to be less than 1 μs, an est. that agrees well with previous exptl. findings of ours for other alkyl peroxynitrites. A comparison of aq. and gaseous thermochem. of alkyl peroxynitrates reveals that alkyl peroxyl radicals and the corresponding alkyl peroxynitrates are similarly solvated by water.
- 29Butkovskaya, N.; Kukui, A.; Le Bras, G. Pressure and Temperature Dependence of Ethyl Nitrate Formation in the C2H5O2 + NO Reaction. J. Phys. Chem. A 2010, 114 (2), 956– 964, DOI: 10.1021/jp910003a29Pressure and Temperature Dependence of Ethyl Nitrate Formation in the C2H5O2 + NO ReactionButkovskaya, Nadezhda; Kukui, Alexandre; Le Bras, GeorgesJournal of Physical Chemistry A (2010), 114 (2), 956-964CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The branching ratio β = k1b/k1a for the formation of Et nitrate, C2H5ONO2, in the gas-phase C2H5O2 + NO reaction, C2H5O2 + NO → C2H5O + NO2 (1a), C2H5O2 + NO → C2H5ONO2 (1b), was detd. over the pressure and temp. ranges 100-600 Torr and 223-298 K, resp., using a turbulent flow reactor (TFR)-ion-mol. reactor(IMR) coupled with a chem. ionization mass spectrometer. At 298 K the C2H5ONO2 yield was found to increase linearly with pressure from about 0.7% at 100 Torr to about 3% at 600 Torr. At each pressure, the branching ratio of C2H5ONO2 formation increases with the decrease of temp. The following parametrization equation has been derived in the pressure and temp. ranges of the study: β(P,T) (%) = (3.88 × 10-3·P (Torr) + 0.365)·(1 + 1500(1/T - 1/298)). The atm. implication of the results obtained is briefly discussed, in particular the impact of β on the evolution of Et nitrate in urban plumes.
- 30Butkovskaya, N. I.; Kukui, A.; Le Bras, G. Pressure Dependence of Iso-Propyl Nitrate Formation in the i-C3H7O2 + NO Reaction. Z. Für Phys. Chem. 2010, 224 (7–8), 1025– 1038, DOI: 10.1524/zpch.2010.6139There is no corresponding record for this reference.
- 31Atkinson, R.; Aschmann, S. M.; Carter, W. P. L.; Winer, A. M.; Pitts, J. N. Alkyl Nitrate Formation from the Nitrogen Oxide (NOx)-Air Photooxidations of C2-C8 n-Alkanes. J. Phys. Chem. 1982, 86 (23), 4563– 4569, DOI: 10.1021/j100220a02231Alkyl nitrate formation from the nitrogen oxide (NOx)-air photooxidations of C2-C8 n-alkanesAtkinson, Roger; Aschmann, Sara M.; Carter, William P. L.; Winer, Arthur M.; Pitts, James N., Jr.Journal of Physical Chemistry (1982), 86 (23), 4563-9CODEN: JPCHAX; ISSN:0022-3654.The yields of alkyl nitrates formed in the NOx-air photooxidns. of the homologous series of n-alkanes from C2H6 [74-84-0] through n-octane [111-65-9] were detd. at 299 ± 2 K and 735 torr total pressure for different chem. systems. Alkyl peroxy radicals were generated by reaction of the n-alkanes with OH radicals (generated from the photolysis of Me nitrite in air) or Cl atoms (from photolysis of Cl in air). The alkyl nitrate yields obtained from the 2 systems, cor. for secondary reactions, were in agreement within the exptl. errors and increased monotonically with the C no. of the n-alkane, from ≤1% for C2H6 to ∼333% for n-octane, with the yields apparently approaching a limit of ∼35% for larger n-alkanes. The relative yields of the various secondary alkyl nitrate isomers in the n-pentane through n-octane systems were in good agreement with those expected from OH radical or Cl atom reaction with the corresponding secondary C-H bonds. However, the relative yields of the primary alkyl nitrates in the C3H8 and C4H10 systems were lower than expected by a factor of ∼2. The data are consistent with the alkyl nitrates being formed almost entirely from the reaction of peroxy radicals with NO, and the ratios of the cor. alkyl nitrate yields thus reflect the fraction of RO2 radicals which react with NO to form alkyl nitrates. These nitrate yields from the reaction of RO2 radicals with NO are important inputs into chem. computer models of the atm. NOx-air photooxidns. of the larger n-alkanes.
- 32Piletic, I. R.; Edney, E. O.; Bartolotti, L. J. Barrierless Reactions with Loose Transition States Govern the Yields and Lifetimes of Organic Nitrates Derived from Isoprene. J. Phys. Chem. A 2017, 121 (43), 8306– 8321, DOI: 10.1021/acs.jpca.7b0822932Barrierless Reactions with Loose Transition States Govern the Yields and Lifetimes of Organic Nitrates Derived from IsoprenePiletic, Ivan R.; Edney, Edward O.; Bartolotti, Libero J.Journal of Physical Chemistry A (2017), 121 (43), 8306-8321CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The chem. reaction mechanism of NO addn. to two β and δ isoprene hydroxy-peroxy radical isomers is examd. in detail using d. functional theory, coupled cluster methods, and the energy resolved master equation formalism to provide ests. of rate consts. and org. nitrate yields. At the M06-2x/aug-cc-pVTZ level, the potential energy surfaces of NO reacting with β-(1,2)-HO-IsopOO• and δ-Z-(1,4)-HO-IsopOO• possess barrierless reactions that produce alkoxy radicals/NO2 and org. nitrates. The nudged elastic band method was used to discover a loosely bound van der Waals (vdW) complex between NO2 and the alkoxy radical that is present in both exit reaction channels. Semiempirical master equation calcns. show that the β org. nitrate yield is 8.5 ± 3.7%. Addnl., a relatively low barrier to C-C bond scission was discovered in the β-vdW complex that leads to direct HONO formation in the gas phase with a yield of 3.1 ± 1.3%. The δ isomer produces a looser vdW complex with a smaller dissocn. barrier and a larger isomerization barrier, giving a 2.4 ± 0.8% org. nitrate yield that is relatively pressure and temp. insensitive. By considering all of these pathways, the first-generation NOx recycling efficiency from isoprene org. nitrates is estd. to be 21% and is expected to increase with decreasing NOx concn.
- 33Aschmann, S. M.; Arey, J.; Atkinson, R. Atmospheric Chemistry of Selected Hydroxycarbonyls. J. Phys. Chem. A 2000, 104 (17), 3998– 4003, DOI: 10.1021/jp993987433Atmospheric Chemistry of Selected HydroxycarbonylsAschmann, Sara M.; Arey, Janet; Atkinson, RogerJournal of Physical Chemistry A (2000), 104 (17), 3998-4003CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Using a relative rate method, rate consts. have been measured at 296 ± 2 K for the gas-phase reactions of the OH radical with 1-hydroxy-2-butanone, 3-hydroxy-2-butanone, 1-hydroxy-3-butanone, 1-hydroxy-2-methyl-3-butanone, 3-hydroxy-3-methyl-2-butanone, and 4-hydroxy-3-hexanone, with rate consts. (in units of 10-12 cm3 mol.-1 s-1) of 7.7 ± 1.7, 10.3 ± 2.2, 8.1 ± 1.8, 16.2 ± 3.4, 0.94 ± 0.37, and 15.1 ± 3.1, resp., where the error limits include the estd. overall uncertainty in the rate const. for the ref. compd. Rate consts. were also measured for reactions with NO3 radicals and O3. Rate consts. for the NO3 radical reactions (in units of 10-16 cm3 mol.-1 s-1) were 1-hydroxy-2-butanone, <9; 3-hydroxy-2-butanone, 6.5 ± 2.2; 1-hydroxy-3-butanone, <22; 1-hydroxy-2-methyl-3-butanone, <22; 3-hydroxy-3-methyl-2-butanone, <2; and 4-hydroxy-3-hexanone, 12 ± 4, where the error limits include the estd. overall uncertainties in the rate consts. for the ref. compds. No reactions with O3 were obsd., and upper limits to the rate consts. of <1.1 × 10-19 cm3 mol.-1 s-1 were derived for all six hydroxycarbonyls. The dominant tropospheric loss process for the hydroxycarbonyls studied here is calcd. to be by reaction with the OH radical.
- 34Barry, J. T.; Berg, D. J.; Tyler, D. R. Radical Cage Effects: The Prediction of Radical Cage Pair Recombination Efficiencies Using Microviscosity Across a Range of Solvent Types. J. Am. Chem. Soc. 2017, 139 (41), 14399– 14405, DOI: 10.1021/jacs.7b0449934Radical Cage Effects: The Prediction of Radical Cage Pair Recombination Efficiencies Using Microviscosity Across a Range of Solvent TypesBarry, Justin T.; Berg, Daniel J.; Tyler, David R.Journal of the American Chemical Society (2017), 139 (41), 14399-14405CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)This study reports a method for correlating the radical recombination efficiencies (FcP) of geminate radical cage pairs to the properties of the solvent. Although bulk viscosity (macroviscosity) is typically used to predict or interpret radical recombination efficiencies, the work reported here shows that microviscosity is a much better parameter. The use of microviscosity is valid over a range of different solvent system types, including nonpolar, arom., polar, and hydrogen bonding solvents. In addn., the relationship of FcP to microviscosity holds for solvent systems contg. mixts. of these solvent types. The microviscosities of the solvent systems were straightforwardly detd. by measuring the diffusion coeff. of an appropriate probe by NMR DOSY spectroscopy. By using solvent mixts., selective solvation was shown to not affect the correlation between FcP and microviscosity. In addn., neither solvent polarity nor radical rotation affects the correlation between FcP and the microviscosity.
- 35Pryor, W. A.; Squadrito, G. L. The Chemistry of Peroxynitrite: A Product from the Reaction of Nitric Oxide with Superoxide. Am. J. Physiol.-Lung Cell. Mol. Physiol. 1995, 268 (5), L699– L722, DOI: 10.1152/ajplung.1995.268.5.L699There is no corresponding record for this reference.
- 36Miyamoto, H.; Yampolski, Y.; Young, C. L. IUPAC-NIST Solubility Data Series. 103. Oxygen and Ozone in Water, Aqueous Solutions, and Organic Liquids (Supplement to Solubility Data Series Volume 7). J. Phys. Chem. Ref. Data 2014, 43 (3), 033102, DOI: 10.1063/1.4883876There is no corresponding record for this reference.
- 37Sprague, M. K.; Garland, E. R.; Mollner, A. K.; Bloss, C.; Bean, B. D.; Weichman, M. L.; Mertens, L. A.; Okumura, M.; Sander, S. P. Kinetics of n-Butoxy and 2-Pentoxy Isomerization and Detection of Primary Products by Infrared Cavity Ringdown Spectroscopy. J. Phys. Chem. A 2012, 116 (24), 6327– 6340, DOI: 10.1021/jp212136r37Kinetics of n-Butoxy and 2-Pentoxy Isomerization and Detection of Primary Products by Infrared Cavity Ringdown SpectroscopySprague, Matthew K.; Garland, Eva R.; Mollner, Andrew K.; Bloss, Claire; Bean, Brian D.; Weichman, Marissa L.; Mertens, Laura A.; Okumura, Mitchio; Sander, Stanley P.Journal of Physical Chemistry A (2012), 116 (24), 6327-6340CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The primary products of n-butoxy and 2-pentoxy isomerization in the presence and absence of O2 have been detected using pulsed laser photolysis-cavity ring down spectroscopy (PLP-CRDS). Alkoxy radicals n-butoxy and 2-pentoxy were generated by photolysis of alkyl nitrite precursors (Bu nitrite or 2-pentyl nitrite, resp.), and the isomerization products with and without O2 were detected by IR cavity ring down spectroscopy 20 μs after the photolysis. We report the mid-IR OH stretch (ν1) absorption spectra for δ-HO-1-C4H8·, δ-HO-1-C4H8OO·, δ-HO-1-C5H10·, and δ-HO-1-C5H10OO·. The obsd. ν1 bands are similar in position and shape to the related alcs. (n-butanol and 2-pentanol), although the HOROO· absorption is slightly stronger than the HOR· absorption. We detd. the rate of isomerization relative to reaction with O2 for the n-butoxy and 2-pentoxy radicals by measuring the relative ν1 absorbance of HOROO· as a function of [O2]. At 295 K and 670 Torr of N2 or N2/O2, we found rate const. ratios of kisom/kO2 = 1.7 (±0.1) × 1019 cm-3 for n-butoxy and kisom/kO2 = 3.4(±0.4) × 1019 cm-3 for 2-pentoxy (2σ uncertainty). Using currently known rate consts. kO2 , we est. isomerization rates of kisom = 2.4 (±1.2) × 105 s-1 and kisom ≈ 3 × 105 s-1 for n-butoxy and 2-pentoxy radicals, resp., where the uncertainties are primarily due to uncertainties in kO2 . Because isomerization is predicted to be in the high pressure limit at 670 Torr, these relative rates are expected to be the same at atm. pressure. Our results include corrections for prompt isomerization of hot nascent alkoxy radicals as well as reaction with background NO and unimol. alkoxy decompn. We est. prompt isomerization yields under our conditions of 4 ± 2% and 5 ± 2% for n-butoxy and 2-pentoxy formed from photolysis of the alkyl nitrites at 351 nm. Our measured relative rate values are in good agreement with and more precise than previous end-product anal. studies conducted on the n-butoxy and 2-pentoxy systems. We show that reactions typically neglected in the anal. of alkoxy relative kinetics (decompn., recombination with NO, and prompt isomerization) may need to be included to obtain accurate values of kisom/kO2 .
- 38Kamath, D.; Mezyk, S. P.; Minakata, D. Elucidating the Elementary Reaction Pathways and Kinetics of Hydroxyl Radical-Induced Acetone Degradation in Aqueous Phase Advanced Oxidation Processes. Environ. Sci. Technol. 2018, 52 (14), 7763– 7774, DOI: 10.1021/acs.est.8b0058238Elucidating the Elementary Reaction Pathways and Kinetics of Hydroxyl Radical-Induced Acetone Degradation in Aqueous Phase Advanced Oxidation ProcessesKamath, Divya; Mezyk, Stephen P.; Minakata, DaisukeEnvironmental Science & Technology (2018), 52 (14), 7763-7774CODEN: ESTHAG; ISSN:0013-936X. (American Chemical Society)Advanced oxidn. processes (AOPs) that produce highly reactive hydroxyl radicals are promising methods to destroy aq. org. contaminants. Hydroxyl radicals react rapidly and nonselectively with org. contaminants and degrade them into intermediates and transformation byproducts. Past studies have indicated that peroxyl radical reactions are responsible for the formation of many intermediate radicals and transformation byproducts. However, complex peroxyl radical reactions that produce identical transformation products make it difficult to exptl. study the elementary reaction pathways and kinetics. In this study, we used ab initio quantum mech. calcns. to identify the thermodynamically preferable elementary reaction pathways of hydroxyl radical-induced acetone degrdn. by calcg. the free energies of the reaction and predicting the corresponding reaction rate consts. by calcg. the free energies of activation. In addn., we solved the ordinary differential equations for each species participating in the elementary reactions to predict the concn. profiles for acetone and its transformation byproducts in an aq. phase UV/hydrogen peroxide AOP. Our ab initio quantum mech. calcns. found an insignificant contribution of Russell reaction mechanisms of peroxyl radicals, but significant involvement of HO2• in the peroxyl radical reactions. The predicted concn. profiles were compared with expts. in the literature, validating our elementary reaction-based kinetic model.
- 39Elford, P. E.; Roberts, B. P. EPR Studies of the Formation and Transformation of Isomeric Radicals [C3H5O]•. Rearrangement of the Allyloxyl Radical in Non-Aqueous Solution Involving a Formal 1,2-Hydrogen-Atom Shift Promoted by Alcohols. J. Chem. Soc., Perkin Trans. 1996, 2 (11), 2247– 2256, DOI: 10.1039/P29960002247There is no corresponding record for this reference.
- 40Konya, K. G.; Paul, T.; Lin, S.; Lusztyk, J.; Ingold, K. U. Laser Flash Photolysis Studies on the First Superoxide Thermal Source. First Direct Measurements of the Rates of Solvent-Assisted 1,2-Hydrogen Atom Shifts and a Proposed New Mechanism for This Unusual Rearrangement. J. Am. Chem. Soc. 2000, 122 (31), 7518– 7527, DOI: 10.1021/ja993570b40Laser Flash Photolysis Studies on the First Superoxide Thermal Source. First Direct Measurements of the Rates of Solvent-Assisted 1,2-Hydrogen Atom Shifts and a Proposed New Mechanism for This Unusual RearrangementKonya, Klara G.; Paul, Thomas; Lin, Shuqiong; Lusztyk, Janusz; Ingold, K. U.Journal of the American Chemical Society (2000), 122 (31), 7518-7527CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The thermal decompn. of bis(4-carboxybenzyl)hyponitrite (I; SOTS-1) in aerated water under physiol. conditions has previously been shown to give the superoxide radical anion in a yield of 40 mol % (Ingold, K. U.; et al. J. Am. Chem. Soc. 1997, 119, 12364). The abs. kinetics of the elementary reactions involved in the cascade of events leading from the first-formed water-sol. benzyloxyl radical to superoxide have been detd. by laser flash photolysis. On the basis of these kinetics it is concluded that SOTS-1 will be suitable for studies of superoxide-induced oxidative stress in most biol. systems. A water-assisted 1,2-H shift converting benzyloxyl into the benzyl ketyl radical is an important step in the above reaction cascade. The kinetics of the 1,2-H shift assisted by H2O, D2O, and a no. of nucleophilic alcs. have been measured for the first time. These data have led to a proposed new mechanism involving the initial formation of a ketyl radical anion and an oxonium cation which generally collapse to give the neutral ketyl radical as the first observable product on the time scale of our expts. (ca. 80 ns).
- 41Schuchmann, H.-P.; Sonntag, C. v. Methylperoxyl Radicals: A Study of the γ-Radiolysis of Methane in Oxygenated Aqueous Solutions. Z. Für Naturforschung B 1984, 39 (2), 217– 221, DOI: 10.1515/znb-1984-0217There is no corresponding record for this reference.
- 42Schuchmann, H.-P.; von Sonntag, C. Photolysis at 185 nm of Dimethyl Ether in Aqueous Solution: Involvement of the Hydroxymethyl Radical. J. Photochem. 1981, 16 (3), 289– 295, DOI: 10.1016/0047-2670(81)80039-142Photolysis at 185 nm of dimethyl ether in aqueous solution: involvement of the hydroxymethyl radicalSchuchmann, Heinz Peter; Von Sonntag, ClemensJournal of Photochemistry (1981), 16 (4), 289-95CODEN: JPCMAE; ISSN:0047-2670.The photolysis of Me2O in aq. soln. at 185 nm yielded CH4, H2, MeOH, (MeOCH2)2, HCHO, MeOCH2CH2OH, (HOCH2)2, EtOH, and MeOEt. These products are explained by three primary processes (formation of CH3O• + •CH3; CH3OCH2• + •H; and CH2O + CH4), the rearrangement process (CH3O•→•CH2OH) known to be undergone by alkoxyl radicals in aq. soln. and subsequent free-radical reactions. In aq. solns. the quantum yield of primary processes leading to products is smaller by about an order of magnitude than those in cyclohexane solns. or those previously found with similar ethers as pure liqs. This apparently means that water as a solvent has a quenching effect. In aq. solns. there is an excited species which is reactive towards N2O and a proton, leading to the formation of N2 and H2 resp. Free hydrated electrons generated by photoionization do not appear to be involved in these reactions.
- 43von Sonntag, C.; Schuchmann, H.-P. The Elucidation of Peroxyl Radical Reactions in Aqueous Solution with the Help of Radiation-Chemical Methods. Angew. Chem., Int. Ed. Engl. 1991, 30 (10), 1229– 1253, DOI: 10.1002/anie.199112291There is no corresponding record for this reference.
- 44Fernández-Ramos, A.; Zgierski, M. Z. Theoretical Study of the Rate Constants and Kinetic Isotope Effects of the 1,2-Hydrogen-Atom Shift of Methoxyl and Benzyloxyl Radicals Assisted by Water. J. Phys. Chem. A 2002, 106 (44), 10578– 10583, DOI: 10.1021/jp020917f44Theoretical Study of the Rate Constants and Kinetic Isotope Effects of the 1,2-Hydrogen-Atom Shift of Methoxyl and Benzyloxyl Radicals Assisted by WaterFernandez-Ramos, Antonio; Zgierski, Marek Z.Journal of Physical Chemistry A (2002), 106 (44), 10578-10583CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Rate consts. and kinetic isotope effects for the 1,2-H shift of methoxyl and benzyloxyl radicals were studied in the presence of water mols. The electronic structure calcns. were carried out at the UB3LYP/6-31G* level, and the dynamics calcns. were performed using the variational transition state theory with semiclassical multidimensional corrections for tunneling. The study deals with 1:1 and 1:2 radical-water complexes in the gas phase. It was found that water catalyzes these rearrangement reactions by forming a bridge contg. two water mols. The dynamics calcns. show that the methoxyl-water complexes react only very slowly. The rate consts. for 1:2 complexes with the benzyloxyl radical are in relative good agreement with the results of laser flash photolysis. The kinetic isotope effects calcd. using heavy water indicate that tunneling makes an important contribution in the 1:1 complex, whereas the contribution due to vibrations is more important for the 1:2 complexes. In both cases, the kinetic isotope effects are substantial. It is concluded that the 1,2-H shift for both radicals is catalyzed by two water mols. through a mechanism that involves the formation of a preliminary 1:1 complex.
- 45Sander, R. Compilation of Henry’s Law Constants (Version 4.0) for Water as Solvent. Atmospheric Chem. Phys. 2015, 15 (8), 4399– 4981, DOI: 10.5194/acp-15-4399-2015There is no corresponding record for this reference.
- 46Mack, J.; Bolton, J. R. Photochemistry of Nitrite and Nitrate in Aqueous Solution: A Review. J. Photochem. Photobiol. Chem. 1999, 128 (1), 1– 13, DOI: 10.1016/S1010-6030(99)00155-046Photochemistry of nitrite and nitrate in aqueous solution: a reviewMack, John; Bolton, James R.Journal of Photochemistry and Photobiology, A: Chemistry (1999), 128 (1-3), 1-14CODEN: JPPCEJ; ISSN:1010-6030. (Elsevier Science S.A.)It has long been known that the photolysis of nitrite and nitrate solns. results in the formation of OH radicals. The mechanism of NO3- photolysis has been the subject of considerable controversy in the literature, however. This review summarizes the exptl. work on NO2- and NO3- photolysis in the context of recent advances in the understanding of the chem. of the peroxynitrite anion (ONOO-) in biol. expts. ONOO- has been found to play a far more significant role in the overall reaction mechanism of NO3- photolysis than had previously been suspected. Research on NO2- and NO3- photolysis, as a pathway to the destruction of org. contaminants in natural waters, is summarized. The possible impact of NO2- and NO3- on Advanced Oxidn. Technologies (AOTs), in which OH radicals are used to initiate the destruction of hazardous org. pollutants in drinking water and industrial waste streams, is explored. A review with 133 refs.
- 47Méreau, R.; Rayez, M.-T.; Caralp, F.; Rayez, J.-C. Isomerisation Reactions of Alkoxy Radicals: Theoretical Study and Structure–Activity Relationships. Phys. Chem. Chem. Phys. 2003, 5 (21), 4828– 4833, DOI: 10.1039/B307708JThere 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.estlett.4c00473.
Additional details on experimental methods, including a diagram of the experimental setup; additional results, including an analysis of the pH dependence of n-pentyl nitrite hydrolysis, an analysis of tubing effects on the observed kinetics, and results of low-NO gas-phase experiments; a detailed derivation of the kinetic model used for extracting rate constants and branching ratios; a comparison between modeled and measured time profiles (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.