importance of sulfate radical anion formation and chemistry …...instance, pyruvate ions have been...

12
Atmos. Chem. Phys., 18, 2809–2820, 2018 https://doi.org/10.5194/acp-18-2809-2018 © Author(s) 2018. This work is distributed under the Creative Commons Attribution 4.0 License. Importance of sulfate radical anion formation and chemistry in heterogeneous OH oxidation of sodium methyl sulfate, the smallest organosulfate Kai Chung Kwong 1 , Man Mei Chim 1 , James F. Davies 2,a , Kevin R. Wilson 2 , and Man Nin Chan 1,3 1 Earth System Science Programme, Faculty of Science, The Chinese University of Hong Kong, Hong Kong, China 2 Chemical Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, CA, USA 3 The Institute of Environment, Energy and Sustainability, The Chinese University of Hong Kong, Hong Kong, China a now at: Department of Chemistry, University of California Riverside, Riverside, CA, USA Correspondence: Man Nin Chan ([email protected]) Received: 30 September 2017 – Discussion started: 24 October 2017 Revised: 18 January 2018 – Accepted: 22 January 2018 – Published: 27 February 2018 Abstract. Organosulfates are important organosulfur compounds present in atmospheric particles. While the abundance, composition, and formation mechanisms of organosulfates have been extensively investigated, it remains unclear how they transform and evolve throughout their atmospheric lifetime. To acquire a fundamental under- standing of how organosulfates chemically transform in the atmosphere, this work investigates the heterogeneous OH radical-initiated oxidation of sodium methyl sulfate (CH 3 SO 4 Na) droplets, the smallest organosulfate detected in atmospheric particles, using an aerosol flow tube reactor at a high relative humidity (RH) of 85 %. Aerosol mass spectra measured by a soft atmospheric pressure ionization source (direct analysis in real time, DART) coupled with a high-resolution mass spectrometer showed that neither functionalization nor fragmentation products are detected. Instead, the ion signal intensity of the bisulfate ion (HSO 4 - ) has been found to increase significantly after OH oxida- tion. We postulate that sodium methyl sulfate tends to fragment into a formaldehyde (CH 2 O) and a sulfate radical anion (SO 4 ·- ) upon OH oxidation. The formaldehyde is likely partitioned back to the gas phase due to its high volatility. The sulfate radical anion, similar to OH radical, can abstract a hydrogen atom from neighboring sodium methyl sulfate to form the bisulfate ion, contributing to the secondary chemistry. Kinetic measurements show that the heterogeneous OH reaction rate constant, k, is (3.79 ± 0.19) × 10 -13 cm 3 molecule -1 s -1 with an effective OH uptake coefficient, γ eff , of 0.17 ± 0.03. While about 40 % of sodium methyl sulfate is being oxidized at the maximum OH exposure (1.27 × 10 12 molecule cm -3 s), only a 3 % decrease in particle diameter is observed. This can be attributed to a small fraction of particle mass lost via the formation and volatilization of formaldehyde. Overall, we firstly demonstrate that the heterogeneous OH oxidation of an organosulfate can lead to the formation of sulfate radical anion and produce inorganic sulfate. Fragmentation processes and sulfate radical anion chemistry play a key role in determining the compositional evolution of sodium methyl sulfate during heterogeneous OH oxidation. 1 Introduction Organosulfur compounds have been found to contribute a significant mass fraction of atmospheric organic compounds. A maximum organosulfur contribution of 30 % to PM 10 or- ganic mass was estimated at a forest site in Hungary by cal- culating the difference between total sulfur and inorganic sul- fate (Surratt et al., 2008). Using a similar approach, Tolocka and Turpin (2012) estimated that organosulfur compounds contribute up to 5–10 % of the total organic mass in south- eastern United States, while Shakya and Peltier (2013) re- ported that organosulfur compounds account for about 1–2 % of organic carbon in Fairbanks, Alaska. Given their high at- mospheric abundances, it is crucial to understand the compo- Published by Copernicus Publications on behalf of the European Geosciences Union.

Upload: others

Post on 28-Jan-2021

0 views

Category:

Documents


0 download

TRANSCRIPT

  • Atmos. Chem. Phys., 18, 2809–2820, 2018https://doi.org/10.5194/acp-18-2809-2018© Author(s) 2018. This work is distributed underthe Creative Commons Attribution 4.0 License.

    Importance of sulfate radical anion formation and chemistry inheterogeneous OH oxidation of sodium methyl sulfate,the smallest organosulfateKai Chung Kwong1, Man Mei Chim1, James F. Davies2,a, Kevin R. Wilson2, and Man Nin Chan1,31Earth System Science Programme, Faculty of Science, The Chinese University of Hong Kong, Hong Kong, China2Chemical Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, CA, USA3The Institute of Environment, Energy and Sustainability, The Chinese University of Hong Kong, Hong Kong, Chinaanow at: Department of Chemistry, University of California Riverside, Riverside, CA, USA

    Correspondence: Man Nin Chan ([email protected])

    Received: 30 September 2017 – Discussion started: 24 October 2017Revised: 18 January 2018 – Accepted: 22 January 2018 – Published: 27 February 2018

    Abstract. Organosulfates are important organosulfurcompounds present in atmospheric particles. While theabundance, composition, and formation mechanisms oforganosulfates have been extensively investigated, it remainsunclear how they transform and evolve throughout theiratmospheric lifetime. To acquire a fundamental under-standing of how organosulfates chemically transform inthe atmosphere, this work investigates the heterogeneousOH radical-initiated oxidation of sodium methyl sulfate(CH3SO4Na) droplets, the smallest organosulfate detectedin atmospheric particles, using an aerosol flow tube reactorat a high relative humidity (RH) of 85 %. Aerosol massspectra measured by a soft atmospheric pressure ionizationsource (direct analysis in real time, DART) coupled witha high-resolution mass spectrometer showed that neitherfunctionalization nor fragmentation products are detected.Instead, the ion signal intensity of the bisulfate ion (HSO4−)has been found to increase significantly after OH oxida-tion. We postulate that sodium methyl sulfate tends tofragment into a formaldehyde (CH2O) and a sulfate radicalanion (SO4·−) upon OH oxidation. The formaldehyde islikely partitioned back to the gas phase due to its highvolatility. The sulfate radical anion, similar to OH radical,can abstract a hydrogen atom from neighboring sodiummethyl sulfate to form the bisulfate ion, contributing tothe secondary chemistry. Kinetic measurements showthat the heterogeneous OH reaction rate constant, k, is(3.79± 0.19)× 10−13 cm3 molecule−1 s−1 with an effectiveOH uptake coefficient, γeff, of 0.17± 0.03. While about

    40 % of sodium methyl sulfate is being oxidized at themaximum OH exposure (1.27× 1012 molecule cm−3 s), onlya 3 % decrease in particle diameter is observed. This canbe attributed to a small fraction of particle mass lost viathe formation and volatilization of formaldehyde. Overall,we firstly demonstrate that the heterogeneous OH oxidationof an organosulfate can lead to the formation of sulfateradical anion and produce inorganic sulfate. Fragmentationprocesses and sulfate radical anion chemistry play a keyrole in determining the compositional evolution of sodiummethyl sulfate during heterogeneous OH oxidation.

    1 Introduction

    Organosulfur compounds have been found to contribute asignificant mass fraction of atmospheric organic compounds.A maximum organosulfur contribution of 30 % to PM10 or-ganic mass was estimated at a forest site in Hungary by cal-culating the difference between total sulfur and inorganic sul-fate (Surratt et al., 2008). Using a similar approach, Tolockaand Turpin (2012) estimated that organosulfur compoundscontribute up to 5–10 % of the total organic mass in south-eastern United States, while Shakya and Peltier (2013) re-ported that organosulfur compounds account for about 1–2 %of organic carbon in Fairbanks, Alaska. Given their high at-mospheric abundances, it is crucial to understand the compo-

    Published by Copernicus Publications on behalf of the European Geosciences Union.

  • 2810 K. C. Kwong et al.: Heterogeneous OH oxidation of sodium methyl sulfate

    Table 1. Chemical structures, properties, rate constant, and effective OH uptake coefficient of sodium methyl sulfate and sodium ethyl sulfate.

    Compounds Sodium methyl sulfate Sodium ethyl sulfate

    Chemical formula

    Molecular formula CH3SO4Na C2H5SO4Na

    Molecular weight 134.0867 148.1147

    Density (g cm−3) 1.60 1.46

    Mass fraction of solute, mfs, at 85 % RH 0.34 0.38

    Heterogeneous OH rate constant, 3.79± 0.19 4.64± 0.29k (× 10−13 cm3 molecule−1 s−1)

    Effective OH uptake coefficient, γeff 0.17± 0.03 0.19± 0.03

    sition, formation, and transformation of organosulfur com-pounds in the atmosphere.

    Organosulfates have been identified as one of the majororganosulfur compounds. The detection of organosulfates inlaboratory studies has been complemented by a number offield observations, which confirm the presence of organosul-fates in atmospheric particles (Iinuma et al., 2007; Surratt etal., 2007, 2008, 2010; Chan et al., 2010; Froyd et al., 2010;Hawkins et al., 2010; Darer et al., 2011; Frossard et al., 2011;Olson et al., 2011; Stone et al., 2012; Shakya and Peltier,2013; Budisulistiorini et al., 2015; Hettiyadura et al., 2015;Huang et al., 2015; Kuang et al., 2016; Rattanavaraha et al.,2016; Riva et al., 2016). Various possible reaction pathwaysby which organosulfates form have been suggested. For in-stance, Iinuma et al. (2009) and Surratt et al. (2010) showedthat organosulfates can be effectively formed via the reac-tive uptake of gas-phase epoxides, which are formed fromthe photooxidation of various biogenic volatile organic com-pounds (e.g., isoprene, α-pinene, and β-pinene), onto theacidic sulfate seed particles. Rudziński et al. (2009) and Noz-ière et al. (2010) suggested that the reactions between sul-fate radical anion and reaction products of isoprene (e.g.,methyl vinyl ketone and methacrolein) can yield a varietyof organosulfates.

    While the abundance, composition, and formation mecha-nisms have extensively been investigated, there is compara-bly little work understanding how organosulfates chemicallytransform in the atmosphere. Organosulfates are primarilypresent in the particle phase owing to their low volatility(Huang et al., 2015; Estillore et al., 2016). They can contin-uously react with gas-phase oxidants such as hydroxyl (OH)radicals, ozone (O3), and nitrate (NO3) radicals at or nearthe particle surface throughout their atmospheric lifetime.These heterogeneous oxidative processes have been foundto change the size, composition, and physiochemical prop-

    erties of both laboratory-generated organic particles and at-mospheric particles (Rudich et al., 2007; George and Ab-batt, 2010; Kroll et al., 2015). To gain a better understand-ing of how organosulfates chemically transform through het-erogeneous oxidation in the atmosphere, this work inves-tigates the heterogeneous OH radical-initiated oxidation ofsodium methyl sulfate (CH3SO4Na) particles, the smallestorganosulfate detected in atmospheric particles, using anaerosol flow tube reactor at a high relative humidity (RH) of85 %. A soft atmospheric pressure ionization source (directanalysis in real time, DART) coupled with a high-resolutionmass spectrometer was employed to characterize the molec-ular composition of the particles before and after OH oxida-tion in real time. Sodium methyl sulfate is detected in atmo-spheric particles with a concentration of 0.7 and 0.34 ng m−3

    during daytime and nighttime, respectively, in Centreville,Alabama (Hettiyadura et al., 2015). As shown in Table 1, thesimple structure of sodium methyl sulfate allows us to gaina more fundamental understanding of the heterogeneous oxi-dation kinetics and chemistry. The sodium salt of methyl sul-fate could be considered as atmospherically relevant since apositive correlation between sodium ion and organosulfateshas been observed over the coastal areas (Sorooshian et al.,2015; Estillore et al., 2016). The effects of salt (e.g., am-monium and potassium salt) on the heterogeneous oxidativekinetics and chemistry are also of atmospheric significanceand warrant future study.

    2 Experimental method

    An atmospheric pressure aerosol flow tube reactor was usedto investigate the heterogeneous OH oxidation of sodiummethyl sulfate droplets. Detail experimental procedures havebeen described previously (Davies and Wilson, 2015; Chimet al., 2017). Briefly, aqueous droplets were generated by aconstant output atomizer and mixed with humidified nitro-

    Atmos. Chem. Phys., 18, 2809–2820, 2018 www.atmos-chem-phys.net/18/2809/2018/

  • K. C. Kwong et al.: Heterogeneous OH oxidation of sodium methyl sulfate 2811

    gen (N2), oxygen (O2), ozone (O3), and hexane (a gas-phaseOH tracer) before introducing into the reactor. The RH in-side the reactor was maintained at 85 % and at a temperatureof 20 ◦C. Estillore et al. (2016) measured the hygroscopicityof sodium methyl sulfate particles and showed that the par-ticles absorb or desorb water reversibly upon increasing ordecreasing RH. These observations suggest that the particleslikely exist as aqueous droplets over a range of RH (10 to90 %). In our experiments, since the sodium methyl sulfatedroplets are always exposed to a high RH, they are likelyaqueous droplets prior to OH oxidation.

    Sodium methyl sulfate droplets were oxidized inside thereactor by gas-phase OH radicals that were generated by thephotolysis of O3 under ultraviolet light (254 nm) illuminationin the presence of water vapor. The OH concentration wasregulated by changing the O3 concentration and determinedby measuring the decay of hexane using gas chromatographycoupled with a flame ionization detector. The OH exposure,defined as the product of OH concentration, [OH], and theparticle residence time, t , was determined by measuring thedecay of the gas-phase tracer, hexane (Smith et al., 2009).

    OH Exposure=−ln([Hex]/[Hex]0)

    kHex=

    t∫o

    [OH]dt, (1)

    where [Hex] is the hexane concentration leaving the re-actor, [Hex]0 is the initial hexane concentration, and kHexis the second-order rate constant of the gas-phase OH–hexane reaction. The aerosol residence time was determinedto be 1.3 min, and the OH exposure was varied from 0 to1.27× 1012 molecule cm−3 s. The particle stream leaving thereactor was passed through an annular Carulite catalyst de-nuder and an activated charcoal denuder to remove O3 andgas-phase species, respectively.

    A portion of the particle stream was sampled by a scan-ning mobility particle sizer (SMPS) for particle size distribu-tion measurements. The remaining flow was delivered into astainless-steel tube heater, where the particles were vapor-ized at 350–400 ◦C. Sodium methyl sulfate particles wereconfirmed to be fully vaporized upon heating at 300 ◦C orabove by measuring the size distribution of the particles leav-ing the heater with the SMPS in a separate experiment. Theresulting gas-phase species were directed to an ionizationregion, a narrow open space between the DART ionizationsource (IonSense: DART SVP), and the inlet orifice of thehigh-resolution mass spectrometer (Thermo Fisher, Q Exac-tive Orbitrap).

    The details of the DART operation have been describedelsewhere (Cody et al., 2005). The DART ionization sourcewas operated in the negative-ion mode. Helium was cho-sen as the ionizing gas and entered an ionization cham-ber, where a high electric potential of 4 kV was applied.This generates a glow discharge containing ions, electrons,and metastable helium atoms. A potential of 200 V was ap-plied to two electrostatic lenses to remove ions and only the

    metastable helium atoms exited the chamber. The gas streamwas heated to 500 ◦C before leaving the ionization source.The metastable helium atoms are responsible for ionizing thegas-phase species in the ionization region (Chan et al., 2014;Cheng et al., 2015, 2016). For ionic compounds like sodiummethyl sulfate, negative ions can be formed via direct ion-ization in the negative-ion mode (Hajslova et al., 2011); forinstance, pyruvate ions have been detected from ammoniumpyruvate using the DART (Block et al., 2010).

    We have run control experiments to investigate the po-tential volatilization of the parent compound (i.e., sodiummethyl sulfate) and the effect of ozone and UV light on thecomposition of the aerosols before oxidation under the sameexperimental conditions. To investigate the volatilization ofsodium methyl sulfate, we have measured the mass spectrumby filtering out the aerosols, and the parent peak is very small,suggesting there is a very small amount of sodium methylsulfate present in the gas phase. Volatilization and gas-phaseoxidation of sodium methyl sulfate is expected to be not sig-nificant. For the effect of ozone and UV light, we found thatthere is no change in aerosol mass spectra in the presenceof ozone without the UV light, suggesting that the reactionof sodium methyl sulfate with ozone is not significant. Theaerosol mass spectrum is about the same as that obtained inthe absence of ozone with the UV light, suggesting that thephotolysis of sodium methyl sulfate aerosols is not likely tooccur.

    3 Results and discussions

    3.1 Aerosol mass spectra

    Figure 1 shows the aerosol mass spectra before and afteroxidation. Before oxidation (Fig. 1a), there is one majorpeak and some minor background peaks. The largest peak atm/z 111 has a chemical formula of CH3SO4−, which is cor-responding to the negative ion (i.e., anionic form) of sodiummethyl sulfate. Figure 1b shows that the intensity of the par-ent compound decreases after oxidation. At the maximumOH exposure (1.27× 1012 molecule cm−3 s), only one newpeak at m/z 97 evolves, corresponding to the bisulfate ion(HSO4−). As shown in Fig. 2, its intensity increases signif-icantly after oxidation, suggesting that the bisulfate ion islikely generated during the oxidation. Based on the aerosolspeciation data measured at different extents of OH oxida-tion, oxidation kinetics will be determined in Sect. 3.2 andreaction mechanisms will be proposed in Sect. 3.3 to ex-plain the formation of major ions detected in the aerosol massspectra.

    www.atmos-chem-phys.net/18/2809/2018/ Atmos. Chem. Phys., 18, 2809–2820, 2018

  • 2812 K. C. Kwong et al.: Heterogeneous OH oxidation of sodium methyl sulfate

    Figure 1. Aerosol mass spectra of sodium methyl sulfate before(a) and after (b) oxidation.

    Figure 2. The kinetic evolution of HSO4− and SO4− as a functionof OH exposure during the heterogeneous OH oxidation of sodiummethyl sulfate. The small uncertainty in ion intensity measurementfor SO4− is not visualized in the figure.

    3.2 Oxidation kinetics

    The normalized parent decay as a function of OH exposureis shown in Fig. 3 and the OH radical-initiated decay can befitted using an exponential function:

    lnI

    I0=−k [OH] · t, (2)

    Figure 3. The normalized decay of sodium methyl sulfate as a func-tion of OH exposure during the heterogeneous OH oxidation.

    where I is the ion signal at a given OH exposure,I0 is the ion signal before oxidation, k is the second-order heterogeneous rate constant, and [OH] · t is theOH exposure. The exponential k is determined to be(3.79± 0.19)× 10−13 cm3 molecule−1 s−1. Assuming a 24-h average OH concentration of 1.5× 106 molecule cm−3,the lifetime of sodium methyl sulfate against heterogeneousOH oxidation is about 20 days. This timescale is longerthan other removal processes such as wet or dry deposi-tion. Laboratory studies have revealed that primary and sec-ondary organosulfates are stable against hydrolysis underatmospheric relevant aerosol acidities and lifetimes, whiletertiary organosulfates may undergo hydrolysis efficiently(Hu et al., 2011). Since sodium methyl sulfate is a primaryorganosulfate, it is expected to be stable against hydrolysis.These results suggest that sodium methyl sulfate is likelychemically stable over atmospherically relevant timescales.Studying the heterogeneous reactivity of sodium methyl sul-fate towards OH radicals provides a much-needed fundamen-tal understanding of the oxidation kinetics and pathways, andthese data may be applied in the interpretation of the oxi-dation of more complex organosulfates, which may have arange of chemical lifetimes in the atmosphere. The effectiveuptake coefficient, γeff, defined as the fraction of OH colli-sions that yield a reaction, is computed (Davies and Wilson,2015):

    γeff =23D0 ρmfsNAMw cOH

    k, (3)

    where D0 is the mean surface-weighted particle diameter, ρis the aerosol density before oxidation, mfs is the mass frac-tion of solute, NA is Avogadro’s number, Mw is the molecu-lar weight of sodium methyl sulfate, and cOH is the averagespeed of gas-phase OH radicals. The mean surface-weighted

    Atmos. Chem. Phys., 18, 2809–2820, 2018 www.atmos-chem-phys.net/18/2809/2018/

  • K. C. Kwong et al.: Heterogeneous OH oxidation of sodium methyl sulfate 2813

    Figure 4. The surface-weighted particle diameter of sodium methylsulfate as a function of OH exposure during heterogeneous OH ox-idation.

    particle diameter was 218 nm and decreased slightly to211 nm (about 3 % decrease) at the maximum OH exposure(Fig. 4). Before oxidation, the composition of the droplets(i.e., mfs) is derived from the hygroscopicity data reportedby Estillore et al. (2016). The particle growth factor, Gf, de-fined as the ratio of the diameter at different RH to the dryparticle at a reference RH (RH1), is converted into mfs usingthe following equation (Ansari and Pandis, 2000; Peng andChan, 2001):

    Gf =

    (mfsRH,1 ρRH,1mfsRH,2 ρRH,2

    ) 13, (4)

    where mfsRH,i and ρRH,i are the mass fraction of solute andparticle density at a given RH, respectively. It is assumed thatsodium methyl sulfate exists as an anhydrous particle at thereference RH (RH < 10 %) (i.e., mfsRH,1 = 1). The particledensity is estimated using the volume additivity rule withthe density of water and sodium methyl sulfate (1.60 g cm−3,Chemistry Dashboard, 2018) with an uncertainty of 20–30 %. The mfs is computed to be 0.34 at 85 %. Using Eq. (3),the γeff is calculated to be 0.17± 0.03. Although the γeff isless than 1, as will be discussed in Sect. 3.3, secondary reac-tions are likely occurring, leading to the formation and sub-sequent reactions of sulfate radical anions (SO4·−).

    3.3 Reaction mechanisms: OH reaction with sodiummethyl sulfate

    Sodium methyl sulfate tends to dissociate and exist in itsionic form because of its high dissociation constant (pKa=−2.4). As shown in Scheme 1, the oxidation is initiatedby hydrogen abstraction from the methyl group by the OHradical, forming an alkyl radical that quickly reacts with

    an oxygen molecule to form a peroxy radical. Based onthe well-known particle-phase reactions (George and Abbatt,2010), the self-reaction of two peroxy radicals can form acarbonyl functionalization product (CHSO5−) via the Ben-nett and Summers mechanism (Bennett and Summers, 1974)or form both alcohol (CH3SO5−) and carbonyl functional-ization products via the Russell reactions (Russell, 1957).Alternatively, alkoxy radicals can be produced through theperoxy–peroxy radical reactions. Once formed, the alkoxyradical can react with an oxygen molecule to form the car-bonyl functionalization product or abstract a hydrogen atomfrom the neighboring molecules to form the alcohol function-alization product. Furthermore, the alkoxy radical can un-dergo fragmentation to yield a formaldehyde (CH2O) and asulfate radical anion (SO4·−). Alternatively, the decomposi-tion of the alkoxy radical, similar to the OH radical-initiatedoxidation of simple alkyl esters (Sun et al., 2012), couldinvolve the rearrangement of the hydrogen atom from thealkoxy radical carbon to the oxygen on the methoxy group.The decomposition of the C−O bond from the methoxygroup generates a bisulfate ion and a formyl radical (CHO).The subsequent reactions of the formyl radical can yield car-bon monoxide (CO) and HO2. Like formaldehyde, carbonmonoxide is volatile and partitions back to the gas phase.

    As shown in Fig. 1, neither functionalization nor fragmen-tation products are detected. Formaldehyde has a mass whichis below the mass range of the mass spectrometer and, onceformed, it is likely partitioned back to the gas phase due toits high volatility. On the other hand, it is expected that thetwo functionalization products can be detected by the DARTionization source if they were formed in significant amounts.Additional experiments were performed to verify whether thealcohol and carbonyl functionalization products can be de-tected by the DART ionization source. We have measuredthe heterogeneous OH radical-initiated oxidation of sodiumethyl sulfate (C2H5SO4Na) under similar experimental con-ditions. As shown in Fig. 5, the negative ions of the alco-hol (C2H5SO5−) and carbonyl (C2H3SO5−) functionaliza-tion products are detected in the aerosol mass spectra. Whenthe sodium ethyl sulfate is oxidized (Fig. 6), the abundance ofthese two functionalization products increases with increas-ing OH exposure (Fig. 7). Similar to sodium methyl sulfate,the bisulfate ion has been detected and its intensity increasesafter oxidation (Fig. 8). These results suggest that, if func-tionalization products are formed during the OH oxidation ofsodium methyl sulfate, they could be detected by the DARTionization source.

    The absence of the functionalization products in theaerosol mass spectra suggests that the OH reaction withsodium methyl sulfate tends to undergo fragmentation pro-cesses rather than functionalization processes. One possibil-ity is that, due to the presence of the bulky sulfate grouprelative to the methyl group, reaction intermediates resultedfrom the self-reaction of two peroxy radicals may not be eas-ily arranged into appropriate configurations (i.e., cyclic tran-

    www.atmos-chem-phys.net/18/2809/2018/ Atmos. Chem. Phys., 18, 2809–2820, 2018

  • 2814 K. C. Kwong et al.: Heterogeneous OH oxidation of sodium methyl sulfate

    Figure 5. Aerosol mass spectra of sodium ethyl sulfate before(a) and after (b) oxidation.

    Figure 6. The normalized parent decay of sodium ethyl sulfate as afunction of OH exposure in the heterogeneous OH oxidation.

    sition states), which are required for the formation of thefunctionalization products via the Russell reaction or Ben-nett and Summers mechanism. Alternatively, alkoxy radicalsare more likely formed, followed by fragmentation. Althoughthe dissociation energies for C–O and C–C bonds of sodiummethyl sulfate are not known, fragmentation processes couldbe enhanced since the decomposition of the alkoxy radicalinvolves the cleavage of a C–O bond, which is in generalthought to be weaker than a C–C bond (Dean and Lang,1992). The bond dissociation energy of a C–O bond might

    Figure 7. The kinetic evolution of carbonyl (C2H3SO5−) and al-cohol (C2H5SO5−) functionalization products as a function ofOH exposure in the heterogeneous OH oxidation of sodium ethylsulfate. The small uncertainty in ion intensity measurement forC2H5SO5− is not visualized in the figure.

    Figure 8. The kinetic evolution of HSO4− and SO4− as a func-tion of OH exposure in the heterogeneous OH oxidation of sodiumethyl sulfate. The small uncertainty in ion intensity measurementfor SO4− is not visualized in the figure.

    be lowered in the presence of a sulfur atom or sulfur–oxygenbearing group (Oae and Doi, 1991; Dean and Lange, 1992).One possibility is that a sulfur atom or sulfur–oxygen bear-ing group (e.g., sulfate) is more electronegative than a carbonatom, reducing the electron density and bond strength of theC–O bond by the inductive effect. Further investigation is re-quired to better understand the effect of sulfate group on thedissociation energies of the C–O bond for the sodium methylsulfate.

    Atmos. Chem. Phys., 18, 2809–2820, 2018 www.atmos-chem-phys.net/18/2809/2018/

  • K. C. Kwong et al.: Heterogeneous OH oxidation of sodium methyl sulfate 2815

    Scheme 1. Proposed reaction mechanism for heterogeneous OH oxidation of sodium methyl sulfate (gray arrows denote the minor pathways).

    3.4 Formation and reaction of sulfate radical anion inthe OH reaction with sodium methyl sulfate

    Scheme 1 shows that sulfate radical anion (SO4·−) can beformed via the fragmentation processes. The sulfate radi-cal anion is a strong oxidant in aqueous phase. For exam-ple, Huie and Clifton (1989) have reported that hydrogen ab-straction by sulfate radical anions on the alkane can resultin the formation of bisulfate ions. They also reported thatthe hydrogen abstraction rate is the highest on the tertiarycarbon, and the rate is 1 order of magnitude smaller for thesecondary carbon, and even smaller for the primary carbon.The second-order rate constants for SO4·− reactions with al-cohols, ethers, alkanes, and aromatic compounds typicallyrange in value from 106 to 109 M−1 s−1 (Neta et al., 1977,1988; Clifton and Huie, 1989; Padmaja et al., 1993). With anaqueous-phase SO4·− concentration of 10−14 M (Herrmannet al., 2000), the calculated lifetime toward aqueous-phaseoxidation with SO4·− ranges from 1.2 days to 3 years. Theseresults suggest that some organic compounds (e.g., alkanesand alkenes) are stable against SO4·−-initiated reactions, butsome (e.g., alcohols and ethers) can react with SO4·− ef-ficiently. Future works are needed to better understand the

    role of SO4·−-initiated oxidation chemistry in the chemicaltransformation of sodium methyl sulfate and organic com-pounds in the atmospheric aerosols. For the OH reaction withsodium methyl sulfate, it is proposed that sulfate radical an-ion, once formed, can abstract a hydrogen atom from theneighboring, unreacted sodium methyl sulfate, yielding thebisulfate ion, which has a small acid dissociation constantin equilibrium with sulfate (SO2−4 ) and hydrogen (H

    +) ions(Ka= 1.2× 10−2) (Brown et al., 2012).

    SO4·−+CH3SO4−→ CH2SO·−4 +HSO4−, (5)

    HSO4− H++SO42−. (6)

    Moreover, the sulfate radical anion may react with particle-phase water to form a bisulfate ion and an OH radical (Tanget al., 1988).

    SO4·−+H2O OH·+HSO4− (7)

    As shown in Fig. 1b, the bisulfate ion is the second largestpeak detected in the aerosol mass spectrum and its intensityhas been found to increase significantly with increasing OHexposure (Fig. 2). The detection of the bisulfate ion providesindirect evidence to support the formation and subsequent

    www.atmos-chem-phys.net/18/2809/2018/ Atmos. Chem. Phys., 18, 2809–2820, 2018

  • 2816 K. C. Kwong et al.: Heterogeneous OH oxidation of sodium methyl sulfate

    Scheme 2. Proposed reaction mechanism of heterogeneous OH oxidation of sodium ethyl sulfate (gray arrows denote the minor pathways).

    reactions of sulfate radical ions. When these reactions oc-cur (Eqs. 5 and 7), additional sodium methyl sulfate is con-sumed by the sulfate radical ions and OH radicals, contribut-ing to the secondary chemistry. It is also known that the self-reaction of two sulfate radical anions can yield a peroxydisul-fate ion (S2O2−8 ) (Hayon et al., 1972; Tang et al., 1988; Huieet al., 1989, 1991; Das, 2001):

    SO4·−+SO4·−→ S2O2−8 . (8)

    Based on its mass-to-charge ratio, the peroxydisulfate ion canbe detected as SO4− at m/z 96 in the aerosol mass spectra(Fig. 1). It is worth noting that the peak at m/z 96 does notlikely originate from the sulfate radical anions due to its highreactivity. The ion signal intensity of the SO4− is measuredto be smaller than that of the bisulfate ion (Fig. 2). However,the abundance of these two ions cannot be directly inferredfrom their intensities owing to their unknown ionization effi-ciencies in the DART ionization source.

    3.5 Sodium methyl sulfate vs. sodium ethyl sulfate:kinetics and chemistry

    We here further examine the results of sodium methyl sul-fate and sodium ethyl sulfate to gain more insights into howthe carbon number affects the kinetics and chemistry for

    these two small organosulfates (C1 and C2). Kinetic mea-surements show that the heterogeneous rate constant and ef-fective OH uptake coefficient of sodium ethyl sulfate aredetermined to be (4.64± 0.29)× 10−13 cm3 molecule−1 s−1

    and 0.19± 0.03, respectively (Table 1 and Fig. 6). Thesekinetic parameters are slightly larger than that of sodiummethyl sulfate (3.79± 0.19× 10−13 cm3 molecule−1 s−1 and0.17± 0.03). An additional carbon atom does not signifi-cantly change the heterogeneous OH reactivity. On the otherhand, the composition of the sodium ethyl sulfate (Fig. 5) isdifferent from that of sodium methyl sulfate after oxidation(Fig. 1). As discussed in Sect. 3.3, the bisulfate ion (HSO4−)and SO4− have been observed for both organosulfates. How-ever, the alcohol and carbonyl functionalization products areonly detected in the OH oxidation of sodium ethyl sulfate.These observations suggest the potential reaction pathwaysmay change with an increasing carbon number.

    As shown in Scheme 2, at the first OH oxidation step ofsodium ethyl sulfate, the hydrogen abstraction can occur ei-ther on the primary (Scheme 2, path A) or the secondary car-bon site (Scheme 2, path B). Depending on the initial OHreaction site, two structural isomers of alcohol (C2H5SO5−)and carbonyl (C2H3SO5−) functionalization products can beformed. However, these isomers cannot be differentiated byexact mass measurements. Although the preferential OH re-

    Atmos. Chem. Phys., 18, 2809–2820, 2018 www.atmos-chem-phys.net/18/2809/2018/

  • K. C. Kwong et al.: Heterogeneous OH oxidation of sodium methyl sulfate 2817

    Figure 9. The surface-weighted particle diameter of sodium ethylsulfate as a function of OH exposure during heterogeneous OH ox-idation.

    action site is not well understood, we postulate that the for-mation of the alcohol and carbonyl functionalization prod-ucts are likely originated from the hydrogen abstraction oc-curred at the primary carbon (Scheme 2, path A). One likelyexplanation is that based on the knowledge of the OH reac-tion with sodium methyl sulfate (Scheme 1), when the hy-drogen abstraction occurs at a carbon atom adjacent to thesulfate group, an alkoxy radical is likely formed from theself-reaction of two peroxy radicals and tends to decompose.It is hypothesized that when the hydrogen atom of the sec-ondary carbon is abstracted by the OH radical (Scheme 2,path B), similar to the sodium methyl sulfate, an alkoxy rad-ical is likely generated and fragments into a sulfate radicalanion and an acetaldehyde, which is volatile and likely par-titions back to the gas phase. The sulfate radical anion cansubsequently react with an unreacted sodium ethyl sulfate,leading to the formation of a bisulfate ion (HSO4−, m/z 97).Alternatively, the self-reactions of two sulfate radical anionscan yield a peroxydisulfate ion (S2O82−), which can be de-tected as SO4− at m/z 96 in the aerosol mass spectra. Futureworks are needed to verify these hypotheses.

    3.6 Aerosol mass lost via volatilization

    While the fragmentation and volatilization processes arelikely the dominant reaction pathways of OH oxidation ofsodium methyl sulfate, the diameter of the particles decreasesslightly from 218 to 211 nm at the maximum OH expo-sure (Fig. 4). As shown in Scheme 1, when the fragmenta-tion processes occur, one methyl group is lost via volatiliza-tion in the form of formaldehyde. The methyl group (CH3,Mw = 15 g mol−1) contributes about 11 % of the total molec-ular mass (CH3SO4Na, Mw = 134 g mol−1). At the maxi-

    mum OH exposure, about 40 % of sodium methyl sulfate isreacted (Fig. 3). If we assume that only fragmentation pro-cesses occur during OH oxidation, this will lead to a 4.4 %loss in particle mass via volatilization. The result of this sim-ple analysis is consistent with the experimental observationthat only a small decrease in particle size (∼ 3 %) is measuredafter oxidation. For the sodium ethyl sulfate (Fig. 9), the par-ticle diameter decreases slightly from 203 to 195.5 nm at themaximum OH exposure (∼ 4 % decrease in particle diame-ter). According to Scheme 2, formaldehyde and acetaldehydeare the volatile fragmentation products, which are likely par-titioned back to the gas phase. Following the above analy-sis, if we assume the fragmentation only leads to the forma-tion and volatilization of the acetaldehyde, this will lead to amaximum 9 % loss in the particle mass at the highest OH ex-posure. Similar to sodium methyl sulfate, the formation andvolatilization of fragmentation products do not cause a sig-nificant decrease in particle mass (and diameter) during OHoxidation.

    4 Conclusions and atmospheric implications

    This work investigates the heterogeneous OH oxidation ofsodium methyl sulfate, the smallest organosulfate found inatmospheric particles. During oxidation, sodium methyl sul-fate tends to fragment into a formaldehyde and a sulfate rad-ical anion. The formation and chemistry of sulfate radicalanions in the heterogeneous OH oxidation of organosulfatescould be of atmospheric interest. This is because sulfate rad-ical anion, like OH radical, can abstract a hydrogen atomfrom unreacted sodium methyl sulfate, contributing to thesecondary reactions. The formation of bisulfate and likelysulfate ions from sulfate radical anion reactions suggests thatOH reaction with sodium methyl sulfate or other organosul-fates can possibly lead to the formation of inorganic sulfate.Moreover, sulfate radical anions can react with organic com-pounds to regenerate organosulfates. Compared to sodiummethyl sulfate, the OH reaction with sodium ethyl sulfateoccurs at a similar reaction rate. The oxidation of both com-pounds can lead to the formation of bisulfate ions, but a dif-ferent distribution of reaction products is observed. Theseobservations suggest that the carbon number plays a signifi-cant role in governing the reaction mechanisms for these twosmall organosulfates. Given a variety of organosulfates havebeen detected in atmospheric particles, the role of molecu-lar structure (e.g., carbon chain length, position, and natureof functional groups, such as hydroxyl and carbonyl) in theheterogeneous OH oxidation kinetics, chemistry, and sulfateradical anion formation and reactions of organosulfates re-mains unexplored and warrants further study.

    Data availability. The underlying research data are available uponrequest from the corresponding author ([email protected]).

    www.atmos-chem-phys.net/18/2809/2018/ Atmos. Chem. Phys., 18, 2809–2820, 2018

  • 2818 K. C. Kwong et al.: Heterogeneous OH oxidation of sodium methyl sulfate

    Competing interests. The authors declare that they have no conflictof interest.

    Acknowledgements. Kai Chung Kwong, Man Mei Chim, andMan Nin Chan are supported by the Direct Grant for Research(4053159), The Chinese University of Hong Kong and HongKong Research Grants Council (HKRGC) project ID 2191111 (ref24300516). James F. Davies and Kevin R. Wilson are supported bythe Director, Office of Energy Research, Office of Basic EnergySciences, Chemical Sciences, Geosciences, and BiosciencesDivision, Condensed Phase and Interfacial Molecular Scienceprogram of the US Department of Energy under contract no.DE-AC02-05CH11231.

    Edited by: Alexander LaskinReviewed by: three anonymous referees

    References

    Ansari, A. S. and Pandis, S. N.: Water Absorption by SecondaryOrganic Aerosol and Its Effect on Inorganic Aerosol Behavior,Environ. Sci. Technol., 34, 71–77. 2000.

    Bennett, J. E. and Summers, R.: Product Studies of the Mutual Ter-mination Reactions of sec-Alkylperoxy Radicals: Evidence forNon-Cyclic Termination, Can. J. Chem., 52, 1377–1379, 1974.

    Block, E., Dane, A. J., Thomas, S., and Cody, R. B.: Applicationsof Direct Analysis in Real Time Mass Spectrometry (DART-MS)in Allium Chemistry. 2-Propenesulfenic and 2-PropenesulfinicAcids, Diallyl Trisulfane S-Oxide, and Other Reactive SulfurCompounds from Crushed Garlic and Other Alliums, J. Agr.Food Chem., 58, 4617–4625, 2010.

    Brown, T. E., LeMay, H. E., Bursten, B. E., Murphy, C. J., andWoodward, P. M.: Chemistry: The Central Science, Pearson Ed-ucation, USA, 12th Edn., Appendix D, 1062–1063, 2012.

    Budisulistiorini, S. H., Li, X., Bairai, S. T., Renfro, J., Liu, Y.,Liu, Y. J., McKinney, K. A., Martin, S. T., McNeill, V. F., Pye,H. O. T., Nenes, A., Neff, M. E., Stone, E. A., Mueller, S.,Knote, C., Shaw, S. L., Zhang, Z., Gold, A., and Surratt, J. D.:Examining the effects of anthropogenic emissions on isoprene-derived secondary organic aerosol formation during the 2013Southern Oxidant and Aerosol Study (SOAS) at the Look Rock,Tennessee ground site, Atmos. Chem. Phys., 15, 8871–8888,https://doi.org/10.5194/acp-15-8871-2015, 2015.

    Chan, M. N., Surratt. J. D., Claeys, M., Edgerton, E. S., Tanner, R.L., Shaw, S. L., Zheng, M., Knipping, E. M., Eddingsaas, N. C.,Wennberg, P. O., and Seinfeld, J. H.: Characterization and Quan-tification of Isoprene-Derived Epoxydiols in Ambient Aerosol inthe Southeastern United States, Environ. Sci. Technol., 44, 4590–4596, 2010.

    Chan, M. N., Zhang, H., Goldstein, A. H., and Wilson, K. R.: TheRole of Water and Phase in the Heterogeneous Oxidation of Solidand Aqueous Succinic Acid Aerosol by Hydroxyl Radicals, J.Phys. Chem. C, 118, 28978–28992, 2014.

    Chemistry Dashboard: Sodium Methyl Sulfate, 512-42-5DTXSID0042406, United States Environmental ProtectionAgency, available at: https://comptox.epa.gov/dashboard/

    dsstoxdb/results?utf8=%E2%9C%93&search=sodium+methyl+sulfate, last access: 22February 2018.

    Cheng, C. T., Chan, M. N., and Wilson, K. R.: The Role of AlkoxyRadicals in the Heterogeneous Reaction of Two Structural Iso-mers of Dimethylsuccinic Acid, Phys. Chem. Chem. Phys., 17,25309–25321, 2015.

    Cheng, C. T., Chan, M. N., and Wilson, K. R.: Importance of Uni-molecular HO2 Elimination in the Heterogeneous OH Reactionof Highly Oxygenated Tartaric Acid Aerosol, J. Phys. Chem. A,120, 5887–5896, 2016.

    Chim, M. M., Chow, C. Y., Davies, J. F., and Chan, M. N.: Ef-fects of Relative Humidity and Particle Phase Water on the Het-erogeneous OH Oxidation of 2−Methylglutaric Acid AqueousDroplets, J. Phys. Chem. A, 121, 1666–1674, 2017.

    Clifton, C. L. and Huie, R. E.: Rate Constant for Hydrogen Ab-straction Reactions of the Sulfate Radical, SO4−, Alcohol, Int. J.Chem. Kinet., 21, 677–687, 1989.

    Cody, R. B., Laramee, J. A., and Durst, H. D.: Versatile New IonSource for the Analysis of Materials in Open Air under AmbientConditions, Anal. Chem., 77, 2297–2302, 2005.

    Darer, A. I., Cole-Filipiak, N. C., O’Connor, A. E., and Elrod, M. J.:Formation and Stability of Atmospherically Relevant Isoprene-Derived Organosulfates and Organonitrates, Environ. Sci. Tech-nol., 45, 1895–1902, 2011.

    Das, T. N.: Reactivity and Role of SO5·− Radical in AqueousMedium Chain Oxidation of Sulfite to Sulfate and AtmosphericSulfuric Acid Generation, J. Phys. Chem. A, 105, 9142–9155,2001.

    Davies, J. F. and Wilson, K. R.: Nanoscale Interfacial GradientsFormed by the Reactive Uptake of OH Radicals onto ViscousAerosol Surfaces, Chem. Sci., 6, 7020–7027, 2015.

    Dean, J. and Lange, N.: Lange’s handbook of chemistry, 15th Edn.,McGraw-Hill, New York, USA, 4.42–4.43, 1992.

    Estillore, A. D., Hettiyadura, A. P. S., and Qin, Z., Leckrone, E.,Wombacher, B., Humphry, T., Stone, E. A., Grassian, V. H.: Wa-ter Uptake and Hygroscopic Growth of Organosulfate Aerosol,Environ. Sci. Technol., 50, 4259–4268, 2016.

    Frossard, A. A., Shaw, P. M., Russell, L. M., Kroll, J. H.,Canagaratna, M. R., Worsnop, D. R., Quinn, P. K., andBates, T. S.: Springtime Arctic Haze Contributions ofSubmicron Organic Particles from European and AsianCombustion Sources, J. Geophys. Res., 116, D05205,https://doi.org/10.1029/2010JD015178, 2011.

    Froyd, K. D., Murphy, S. M., Mruphy, D. M., de Gouw, J. A., Ed-dingsaas, N. C., and Wennberg, P. O.: Contribution of Isoprene-Derived Organosulfates to Free Tropospheric Aerosol Mass, P.Natl. Acad. Sci. USA, 107, 21360–21365, 2010.

    George, I. J. and Abbatt, J. P. D.: Heterogeneous Oxidation of At-mospheric Aerosol Particles by Gas-Phase Radials, Nat. Chem.,2, 713–722, 2010.

    Hajslova, J., Cajka, T., and Vaclavik, L.: Challenging Applica-tions Offered by Direct Analysis in Real Time (DART) in Food-Quality and Safety Analysis, TRAC-Trends Anal. Chem., 30,204–218, 2011.

    Hawkins, L. N., Russell, L. M., Covert, D. S., Quinn, P. K., andBates, T. S.: Carboxylic Acids, Sulfates, and Organosulfates inProcessed Continental Organic Aerosol over the Southeast Pa-cific Ocean during VOCALS-REx 2008, J. Geophys. Res., 115,D13201, https://doi.org/10.1029/2009JD013276, 2010.

    Atmos. Chem. Phys., 18, 2809–2820, 2018 www.atmos-chem-phys.net/18/2809/2018/

    https://doi.org/10.5194/acp-15-8871-2015https://comptox.epa.gov/dashboard/dsstoxdb/results?utf8=%E2%9C%93&search=sodium+methyl+sulfatehttps://comptox.epa.gov/dashboard/dsstoxdb/results?utf8=%E2%9C%93&search=sodium+methyl+sulfatehttps://comptox.epa.gov/dashboard/dsstoxdb/results?utf8=%E2%9C%93&search=sodium+methyl+sulfatehttps://doi.org/10.1029/2010JD015178https://doi.org/10.1029/2009JD013276

  • K. C. Kwong et al.: Heterogeneous OH oxidation of sodium methyl sulfate 2819

    Hayon, E., Treinin, A., and Wilf, J.: Electronic Spectra, Photo-chemistry, and Autoxidation Mechanism of the Sulfite-Bisulfite-Pyrosulfite System. The SO2−, SO3−, SO4− and SO5− Radi-cals, J. Am. Chem. Soc., 94, 47–57, 1972.

    Herrmann, H., Ervens, B., Jacobi, H. W., Wolke, R., Nowacki, P.,and Zellner, R.: CAPRAM2.3: A Chemical Aqueous Phase Rad-ical Mechanism for Tropospheric Chemistry, J. Atmos. Chem.,36, 231–284, 2000.

    Hettiyadura, A. P. S., Stone, E. A., Kundu, S., Baker, Z., Geddes,E., Richards, K., and Humphry, T.: Determination of atmosphericorganosulfates using HILIC chromatography with MS detection,Atmos. Meas. Tech., 8, 2347–2358, https://doi.org/10.5194/amt-8-2347-2015, 2015.

    Hu, K. S., Darer, A. I., and Elrod, M. J.: Thermodynamics andkinetics of the hydrolysis of atmospherically relevant organon-itrates and organosulfates, Atmos. Chem. Phys., 11, 8307–8320,https://doi.org/10.5194/acp-11-8307-2011, 2011.

    Huang, D. D., Li, Y. J., Lee, B. P., and Chan, C. K.: Analysisof Organic Sulfur Compounds in Atmospheric Aerosols at theHKUST Supersite in Hong Kong Using HR-ToF-AMS, Environ.Sci. Technol., 49, 3672–3679, 2015.

    Huie, R. E. and Clifton, C. L.: Rate Constants for Hydrogen Ab-straction Reactions of the Sulfate Radical, SO4−, Alkanes andEthers, Int. J. Chem. Kinet., 21, 611–619, 1989.

    Huie, R. E., Clifton, C. L., and Altstein, N.: A Pulse Radiolysis andFlash Photolysis Study of the Radicals SO2−, SO3−, SO4− andSO5−, Radiat. Phys. Chem., 33, 361–370, 1989.

    Huie, R. E., Clifton, C. L., and Neta, P.: Electron Transfer Reac-tion Rates and Equilibria of the Carbonate and Sulfate RadicalAnions, Radiat. Phys. Chem., 38, 477–481, 1991.

    Iinuma, Y., Müller, C., Berndt, T., Böge, O., Claeys, M., and Her-rmann, H.: Evidence for the Existence of Organosulfates from β-pinene Ozonolysis in Ambient Secondary Organic Aerosol, En-viron. Sci. Technol., 41, 6678–6683, 2007.

    Iinuma, Y., Böge, O., Kahnt, A., and Herrmann, H.: LaboratoryChamber Studies on the Formation of Organosulfates from Reac-tive Uptake of Monoterpene Oxides, Phys. Chem. Chem. Phys.,11, 7985–7997, 2009.

    Kroll, J. H., Lim, C. Y., Kessler, S. H., and Wilson, K. R.: Hetero-geneous Oxidation of Atmospheric Organic Aerosol: Kinetics ofChanges to the Amount and Oxidation State of Particle-PhaseOrganic Carbon, J. Phys. Chem. A, 119, 10767–10783, 2015.

    Kuang, B. Y., Lin, P., Hu, M., and Yu, J. Z.: Aerosol Size Distri-bution Characteristics of Organosulfates in the Peral River DeltaRegion, China, Atmos. Environ., 130, 23–35, 2016.

    Neta, P., Madhavan, V., Zamel, H., and Fessenden, R. W.: RateConstants and Mechanism of Reaction of SO4·− with AromaticCompounds, J. Am. Chem. Soc., 99, 163–164, 1977.

    Neta, P., Huie, R. E., and Ross, A. B.: Rate Constants for Reactionsof Inorganic Radicals in Aqueous Solution, J. Phys. Chem. Ref.Data, 17, 1027–1284, https://doi.org/10.1063/1.555808, 1988.

    Nozière, B., Ekström, S., Alsberg, T., and Holmström, S.:RadicalInitiated Formation of Organosulfates and Surfactantsin Atmospheric Aerosols, Geophys. Res. Lett., 37, L05806,https://doi.org/10.1029/2009GL041683, 2010.

    Oae, S. and Doi, J.: Organic Sulfur Chemistry: Structure and Mech-anism, CRC Press, Boca Raton, FL, USA, 1991.

    Olson, C. N., Galloway, M. M., Yu, G., Hedman, C. J., Lockett,M. R., Yoon, T., Stone, E. A., Smith, L. M., and Keutsch, F.

    N.: Hydroxycarboxylic Acid-Derived Organosulfates: Synthesis,Stability, and Quantification in Ambient Aerosol, Environ. Sci.Technol., 45, 6468–6474, 2011.

    Padmaja, S., Alfassi, Z. B., Neta, P., and Huie, R. E.: Rate Constantsfor Reactions of SO4·− Radicals in Acetonitrile, Int. J. Chem.Kinet., 25, 193–198, 1993.

    Peng, C. and Chan, C. K.: The Water Cycle of Water-Soluble Or-ganic Salts of Atmospheric Importance, Atmos. Environ., 35,1183–1192, 2001.

    Rattanavaraha, W., Chu, K., Budisulistiorini, S. H., Riva, M., Lin,Y.-H., Edgerton, E. S., Baumann, K., Shaw, S. L., Guo, H.,King, L., Weber, R. J., Neff, M. E., Stone, E. A., Offenberg, J.H., Zhang, Z., Gold, A., and Surratt, J. D.: Assessing the im-pact of anthropogenic pollution on isoprene-derived secondaryorganic aerosol formation in PM2.5 collected from the Birm-ingham, Alabama, ground site during the 2013 Southern Oxi-dant and Aerosol Study, Atmos. Chem. Phys., 16, 4897–4914,https://doi.org/10.5194/acp-16-4897-2016, 2016.

    Riva, M., Da Silva Barbosa, T., Lin, Y.-H., Stone, E. A., Gold,A., and Surratt, J. D.: Chemical characterization of organosul-fates in secondary organic aerosol derived from the photoox-idation of alkanes, Atmos. Chem. Phys., 16, 11001–11018,https://doi.org/10.5194/acp-16-11001-2016, 2016.

    Rudich, Y., Donahue, N. M., and Mentel, T. F.: Aging of OrganicAerosol: Bridging the Gap Between Laboratory and Field Stud-ies, Annu. Rev. Phys. Chem., 58, 321–352, 2007.

    Rudziński, K. J., Gmachowski, L., and Kuznietsova, I.: Reactions ofisoprene and sulphoxy radical-anions – a possible source of at-mospheric organosulphites and organosulphates, Atmos. Chem.Phys., 9, 2129–2140, https://doi.org/10.5194/acp-9-2129-2009,2009.

    Russell, G. A.: Deuterium-Isotope Effects in the Autoxidation ofAralkyl Hydrocarbons. Mechanism of the Interaction of PeroxyRadicals, J. Am. Chem. Soc., 79, 3871–3877, 1957.

    Shakya, K. M. and Peltier, R. E.: Investigating Missing Sources ofSulfur at Fairbanks, Alaska, Environ. Sci. Technol., 47, 9332–9338, 2013.

    Smith, J. D., Kroll, J. H., Cappa, C. D., Che, D. L., Liu, C. L.,Ahmed, M., Leone, S. R., Worsnop, D. R., and Wilson, K.R.: The heterogeneous reaction of hydroxyl radicals with sub-micron squalane particles: a model system for understanding theoxidative aging of ambient aerosols, Atmos. Chem. Phys., 9,3209–3222, https://doi.org/10.5194/acp-9-3209-2009, 2009.

    Sorooshian, A., Crosbie, E., Maudlin, L. C., Youn, J. S., Wang, Z.,Shingler, T., Ortega, A. M., Hersey, S., Woods, R. K.: Surface andAirborne Measurements of Organosulfur and Methanesulfonateover the Western United States and Coastal Areas, J. Geophys.Res.-Atmos., 120, 8535–8548, 2015.

    Stone, E. A., Yang, L., Yu, L. E., and Rupakheti, M.: Characteri-zation of Organosulfates in Atmospheric Aerosols at Four AsianLocations, Atmos. Environ., 47, 323–329, 2012.

    Sun, X., Hu, Y., Xu, F., Zhang, Q., and Wang, W.: Mechanism andKinetic Studies for OH Radical-initiated Atmospheric Oxidationof Methyl Propionate, Atmos. Environ., 63, 14–21, 2012.

    Surratt, J. D., Kroll, J. H., Kleindienst T. E., Edney, E. O., Claeys,M., Sorooshian, A., Ng, N. L., Offenberg, J. H., Lewandowski,M., Jaoui, M., Flagan, R. C., and Seinfeld, J. H.: Evidencefor Organosulfates in Secondary Organic Aerosol, Environ. Sci.Technol., 41, 517–527, 2007.

    www.atmos-chem-phys.net/18/2809/2018/ Atmos. Chem. Phys., 18, 2809–2820, 2018

    https://doi.org/10.5194/amt-8-2347-2015https://doi.org/10.5194/amt-8-2347-2015https://doi.org/10.5194/acp-11-8307-2011https://doi.org/10.1063/1.555808https://doi.org/10.1029/2009GL041683https://doi.org/10.5194/acp-16-4897-2016https://doi.org/10.5194/acp-16-11001-2016https://doi.org/10.5194/acp-9-2129-2009https://doi.org/10.5194/acp-9-3209-2009

  • 2820 K. C. Kwong et al.: Heterogeneous OH oxidation of sodium methyl sulfate

    Surratt, J. D., Gonzaìlez, Y. G., Chan, A. W. H., Vermeylen, R.,Shahgholi, M., Kleindienst, T. E., Edney, E. O., Offenberg, J. H.,Lewandowski, M., Jaoui, M., Maenhaut, W., Claeys, M., Flagan,R. C., and Seinfeld, J. H.: Organosulfate Formation in BiogenicSecondary Organic Aerosol, J. Phys. Chem. A, 112, 8345–8378,2008.

    Surratt, J. D., Chan, A. W. H., Eddingsaas, N. C., Chan, M. N.,Loza, C. L., Kwan, A. J., Hersey, S. P., Flagan, R. C., Wennberg,P. O., and Seinfeld, J. H.: Reactive Intermediates Revealed inSecondary Organic Aerosol Formation from Isoprene, P. Natl.Acad. Sci. USA, 107, 6640–6645, 2010.

    Tang, Y., Thorn, R. P., Mauldin III, R. L., and Wine, P. H.: Kineticsand Spectroscopy of the SO4− Radical in Aqueous Solution, J.Photochem. Photobio. A, 44, 243–258, 1988.

    Tolocka, M. P. and Turpin, B.: Contribution of Organosulfur Com-pounds to Organic Aerosol Mass, Environ. Sci. Technol., 46,7978–7983, 2012.

    Atmos. Chem. Phys., 18, 2809–2820, 2018 www.atmos-chem-phys.net/18/2809/2018/

    AbstractIntroductionExperimental methodResults and discussionsAerosol mass spectraOxidation kineticsReaction mechanisms: OH reaction with sodium methyl sulfateFormation and reaction of sulfate radical anion in the OH reaction with sodium methyl sulfateSodium methyl sulfate vs. sodium ethyl sulfate: kinetics and chemistryAerosol mass lost via volatilization

    Conclusions and atmospheric implicationsData availabilityCompeting interestsAcknowledgementsReferences