Aqueous Processing of Atmospheric Organic Particles in Cloud Water

Jun 11, 2015 - ABSTRACT: Cloudwater and below-cloud atmospheric par- ticle samples were collected onboard a research aircraft during the Southern Oxid...
0 downloads 5 Views 369KB Size
Subscriber access provided by NEW YORK UNIV

Article

Aqueous Processing of Atmospheric Organic Particles in Cloud Water Collected via Aircraft Sampling Eric J Boone, Alexander Laskin, Julia Laskin, Christopher Wirth, Paul B. Shepson, Brian H. Stirm, and Kerri A. Pratt Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.5b01639 • Publication Date (Web): 11 Jun 2015 Downloaded from http://pubs.acs.org on June 16, 2015

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Environmental Science & Technology is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 31

Environmental Science & Technology

    1 

Aqueous Processing of Atmospheric Organic Particles in Cloud Water



Collected via Aircraft Sampling



Eric J. Boone1, Alexander Laskin2, Julia Laskin3, Christopher Wirth4, Paul B. Shepson4,5,6, Brian



H. Stirm7, Kerri A. Pratt1,8,*

5  6  7  8  9  10  11  12  13  14 

1

Department of Chemistry, University of Michigan, Ann Arbor, MI, USA

2

Environmental Molecular Sciences Laboratory, Pacific Northwest National Laboratory,

Richland, WA USA 3

Physical Sciences Division, Pacific Northwest National Laboratory, Richland, WA, USA

4

Department of Chemistry, Purdue University, West Lafayette, IN, USA

5

Department of Earth, Atmospheric, and Planetary Sciences, Purdue University, West Lafayette,

IN, USA 6

Purdue Climate Change Research Center, Purdue University, West Lafayette, IN, USA

7

Department of Aviation Technology, Purdue University, West Lafayette, IN, USA

8

Department of Earth and Environmental Sciences, University of Michigan, Ann Arbor, MI,

15 

USA

16 

* Corresponding Author: Kerri A. Pratt

17 

Department of Chemistry

18 

University of Michigan

19 

930 N. University Ave.

20 

Ann Arbor, MI 48109 USA

21 

[email protected]

22 

(734) 763-2871

23  1   

ACS Paragon Plus Environment

Environmental Science & Technology

Page 2 of 31

    24 

Abstract

25 

Cloud water and below-cloud atmospheric particle samples were collected onboard a research

26 

aircraft during the Southern Oxidant and Aerosol Study (SOAS) over a forested region of

27 

Alabama in June 2013. The organic molecular composition of the samples was studied to gain

28 

insights into the aqueous-phase processing of organic compounds within cloud droplets. High

29 

resolution mass spectrometry (HRMS) with nanospray desorption electrospray ionization (nano-

30 

DESI) and direct infusion electrospray ionization (ESI) were utilized to compare the organic

31 

composition of the particle and cloud water samples, respectively. Isoprene and monoterpene-

32 

derived organosulfates and oligomers were identified in both the particles and cloud water,

33 

showing the significant influence of biogenic volatile organic compound oxidation above the

34 

forested region. While the average O:C ratios of the organic compounds were similar between

35 

the atmospheric particle and cloud water samples, the chemical composition of these samples

36 

was quite different. Specifically, hydrolysis of organosulfates and formation of nitrogen-

37 

containing compounds were observed for the cloud water when compared to the atmospheric

38 

particle samples, demonstrating that cloud processing changes the composition of organic

39 

aerosol.

2   

ACS Paragon Plus Environment

Page 3 of 31

Environmental Science & Technology

    40 

Introduction

41 

Atmospheric aerosols play a significant role in the Earth’s climate system by scattering

42 

and absorbing light and impacting cloud formation.1 By acting as cloud condensation nuclei

43 

(CCN), organic-containing particles provide a source of organic material to cloud droplets.2 The

44 

composition of these cloud droplets can influence precipitation chemistry, impacting ecosystems;

45 

in addition, if droplet evaporation instead occurs, the residual particle chemical composition may

46 

impact future cloud formation.2 Secondary organic aerosol (SOA), formed through the oxidation

47 

of volatile organic compounds, represents a significant mass fraction of the total submicron

48 

aerosol globally in the atmosphere, but there remain uncertainties in predictions of SOA mass

49 

concentrations.3 Under-prediction of the total organic aerosol burden in the spring and summer in

50 

the mid-latitudes suggests an unaccounted source of organic mass.4 SOA formation via aqueous

51 

reactions of dissolved organic gases and particles in cloud droplets has been suggested as a

52 

potential candidate for the missing source of SOA in models.5 Indeed, inclusion of a

53 

parameterization of SOA formation in cloud droplets through aqueous reactions within a regional

54 

chemical transport model improved agreement with measurements of organic aerosol mass.4

55 

However, our understanding of these processes is limited by a paucity of experimental data on

56 

the molecular characterization of organic compounds in ambient cloud water.5-11 Many laboratory studies have examined the production of high molecular weight organic

57  58 

compounds through reactions in aqueous solutions using atmospherically relevant species.5,

12

59 

These compounds are produced through a variety of pathways, including radical reactions,

60 

inorganic-organic reactions, photochemistry, hydration, and oligomerization.12 Aqueous

61 

oxidation reactions have been proposed to increase the overall (bulk) O:C ratio of organic

62 

compounds in cloud water and SOA.5 However, the dominant reactions occurring in cloud

3   

ACS Paragon Plus Environment

Environmental Science & Technology

Page 4 of 31

    63 

droplets may be different from deliquesced aerosols, which are more concentrated and have

64 

longer lifetimes; these cloud water conditions are expected to inhibit aqueous reactions leading

65 

to oligomer and organosulfate formation.5 Further, organosulfates, produced in deliquesced

66 

aerosol and activated as cloud droplets, could be additionally modified in cloud droplets through

67 

aqueous phase reactions such as hydrolysis.13 In contrast to aqueous oxidation reactions,

68 

hydrolysis reactions of organonitrates and organosulfates would reduce the O:C ratio of the

69 

resulting products.14 Aqueous-phase reactions can also create more volatile compounds through

70 

fragmentation of larger species, resulting in a net organic mass loss.12 These are examples of the

71 

complexity and competing factors in aqueous reactions.

72 

The southeastern United States is a unique ecosystem for studying aqueous-phase

73 

reactions within atmospheric particles and cloud water. The combination of significant biogenic

74 

and anthropogenic emissions in the humid region has been shown to result in a cooling haze,

75 

consisting of aqueous particles.15 Previous studies demonstrated that 70 - 80% of water soluble

76 

organic carbon in atmospheric particles collected near Atlanta, GA originated from biogenic

77 

emissions.16 Modeling of the spatial distribution of water soluble volatile organic compounds

78 

and particulate liquid water content in the eastern United States indicates that anthropogenic

79 

emissions lead to higher particulate mass, which increases the amount of water condensed onto

80 

the particles.17 This may make aqueous processing a dominant SOA production mechanism in

81 

the region.17 During the 2013 Southern Oxidant and Aerosol Study (SOAS) in Alabama, samples

82 

of atmospheric particles and cloud water were collected onboard research aircraft for later

83 

characterization of their organic constituents using high resolution mass spectrometry.

84 

Comparison of the composition of the below-cloud atmospheric particles with the cloud water

85 

provides insights into potential aqueous-phase processing of the organic species. Several hundred

4   

ACS Paragon Plus Environment

Page 5 of 31

Environmental Science & Technology

    86 

unique organic compounds, many derived from biogenic precursors, were identified with

87 

significant differences observed between the particle and cloud water samples, suggesting that

88 

cloud processing plays a significant role in altering the composition of the organic compounds.

89  90 

Experimental

91 

Atmospheric particle and cloud water samples were collected during three SOAS flights

92 

(June 10, 13, and 27, 2013) of a Beechcraft Duchess twin-engine aircraft (Purdue University

93 

Airborne Laboratory for Atmospheric Research). Sampling was conducted between ~13:00 and

94 

~16:00 local time (central daylight time, CDT) over the region ~40km northeast of the SOAS

95 

Centreville site (32.90N, 87.25W). Atmospheric particles were collected below cloud (~1000-

96 

1200 masl) during ~30-45 min of sampling, and cloud water was collected through the middle

97 

and upper sections (~1800-2000 masl) of developing cumulus clouds. During cloud water and

98 

particle collection, convective available potential energy (CAPE) was measured at the Shelby

99 

County

Airport

in

Birmingham,

AL

by

soundings

100 

(http://weather.uwyo.edu/upperair/sounding.html) recorded at 12:00 UTC (19:00 CDT). The

101 

CAPE values were 1069 J/kg, 1951 J/kg, and 2716 J/kg for the flights on June 10, 13, and 27,

102 

respectively. These values designate high atmospheric instability, indicating that the convective

103 

clouds were significantly influenced by the atmospheric boundary layer.

104 

Samples of atmospheric particles were first collected during multiple passages of the

105 

aircraft below cloud using a modified rotating drum impactor.18 Particle samples in three size

106 

ranges (1.15-2.5 μm - stage A, 0.34-1.15 μm - stage B, and 0.07-0.34 μm - stage C) were

107 

collected on pre-baked aluminum foil. Here, we focus on stage B samples, which had sufficient

108 

organic mass for analysis and correspond to particles that are expected have acted as CCN to

5   

ACS Paragon Plus Environment

Environmental Science & Technology

Page 6 of 31

    109 

form cloud droplets. Next, the aircraft made several passages through the overlying clouds with a

110 

modified Mohnen slotted rod cloud water collector19 extended out of the top of the aircraft; cloud

111 

water samples (ranging from 5 - 15 mL each) were collected following the method of Hill et al.20

112 

Cloud droplets were collected into glass vials after impacting the collector Teflon rods, which

113 

are characterized by a 50% cutoff diameter of ~5.5 μm.21 After aircraft landing, particle and

114 

cloud water samples were kept at -40°C pending analysis.

115 

High resolution negative ion mass spectra of the particle samples were obtained using a

116 

Thermo LTQ Orbitrap mass spectrometer equipped with a custom-built nano spray desorption

117 

electrospray ionization (nano-DESI) source described in detail elsewhere.22, 23 Briefly, the nano-

118 

DESI probe is comprised of two fused silica capillaries: one supplying a solvent from a syringe

119 

pump operated at a typical flow rate of ~1µL/min and another capillary removing the solvent

120 

containing the extracted analyte molecules and delivering it to the mass spectrometer inlet. The

121 

resulting mixture is ionized by applying ~4 kV to the first capillary. Charged micro-droplets

122 

produced at the instrument inlet are subsequently desolvated in a heated capillary operated at 250

123 

°C.22, 23 In this study, the instrument was operated at a mass resolution of m/m =100,000 at m/z

124 

400 and a mass range of m/z = 150-2000.23 A 70/30 (volume) mixture of acetonitrile (HPLC

125 

grade, 99.9%) and milli-q (18 MOhm) water was used as the nano-DESI solvent. The solvent

126 

droplet was drawn directly across each particle sample spot on the aluminum substrate for

127 

analysis. A background mass spectrum was acquired over the bare aluminum substrate next to

128 

the particle sample. To assist in ionization and desolvation, the cloud water was diluted to a

129 

mixture of 70/30 (by volume) HPLC grade acetonitrile and cloud water. The resulting solution

130 

was directly injected into the Thermo LTQ Orbitrap mass spectrometer using an ESI source

131 

operated at a typical flow rate of ~0.5 µL/min and a high voltage of ~3.5 kV. The acquired mass

6   

ACS Paragon Plus Environment

Page 7 of 31

Environmental Science & Technology

    132 

range for the ESI experiment was m/z 50-2000 amu. A field blank of ultrapure water sprayed

133 

across the cloud water collector, prior to each cloud water sampling period, was diluted to a

134 

70/30 (by volume) mixture of HPLC grade acetonitrile and blank water; this was analyzed before

135 

and after each cloud water injection.

136 

Mass spectral peaks of over 3 times signal/noise were extracted from the raw files using

137 

the Decon2Ls program (http://omics.pnl.gov/software/decontools-decon2ls) developed at Pacific

138 

Northwest National Laboratory (PNNL), and features with enhancements of less than 5 times of

139 

blank intensity, as well as 13C isotopes, were removed from the final list. The final peak list was

140 

grouped into homologous series using second order mass defect analysis (bases CH2 and H2),

141 

described in Roach et al.24 Molecular formula assignment for representative peaks from each

142 

group was performed using the Formula Calculator v 1.1 from the National High Magnetic Field

143 

Laboratory (http://www.magnet.fsu.edu/usershub/scientificdivisions/icr/icr_software.html). The

144 

parameters used for the assignment were: C0-50, H0-100, O0-20, N0-3, S0-2 in negative mode and C0-

145 

50,

146 

deprotonation exclusively, while positive ions were assumed to be formed by protonation or

147 

addition of sodium. Identification of one peak from each group enabled formula assignments for

148 

all other members of the homologous group. All peaks were assigned with a mass accuracy of ±

149 

3 ppm, and at least 55% of the peaks and 65% of the total signal intensity were assigned

150 

molecular formulas for each sample. Because over 90% of the peaks were below m/z 400 and the

151 

increased uncertainty for assigning large masses, we focused our analysis on m/z < 400. For

152 

number fraction comparisons, only peaks from m/z 150-400 were considered. Bulk O:C, H:C,

153 

and N:C ratios were calculated by averaging the individual atomic ratios of each compound

154 

assigned a formula in a given sample.25 This procedure was also used to calculate average double

H0-100, O0-20, N0-3, S0-1, Na0-1 in positive mode. Negative ions were assumed to have formed via

7   

ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 31

    155 

bond equivalencies (DBE). Possible organosulfate compounds were identified with formulas

156 

having 1 sulfur atom and at least 5 oxygen atoms to account for the sulfate functionality, as well

157 

as other oxygen-containing functional groups that were likely present. Organosulfates were not

158 

expected to form within the ESI source, based on previous results from laboratory studies.6

159  160 

Results and Discussion

161 

The mass spectral features observed in the particle and cloud water samples indicated a

162 

diverse mixture of organic compounds with a variety of functional groups. Example negative ion

163 

mass spectra from the June 10th particle and cloud water samples are shown in Figure 1. Mass

164 

spectra, as well as a complete list of peaks and assignments, for all other samples can be found in

165 

the supplemental information (Figures S1-S2, Tables S1-S2). CHO compounds were the most

166 

abundant by number and intensity, averaging ~50% of assigned peaks by number across all

167 

samples. CHOS componds were the next most prevalent for the particle samples (~30% by

168 

number on average), while CHON compounds were more numerous in the cloud water (~20%

169 

by number on average). CHONS compounds were observed in small amounts in both particle

170 

and cloud spectra, averaging ~8% by number across all samples. These trends are consistent with

171 

previous high resolution mass spectrometry studies of atmospheric particles26,

172 

water.28

27

and cloud

173 

To gain insights into cloud processing of organic compounds, the organic molecular

174 

composition of atmospheric particles below convective cumulus clouds was compared to that of

175 

the cloud water, with the implicit assumptions that particles acted as CCN to form the cloud

176 

droplets and that the composition of the particle samples represents the initial composition of the

177 

cloud water before any aqueous chemical reactions occur. Following cloud droplet activation,

8   

ACS Paragon Plus Environment

Page 9 of 31

Environmental Science & Technology

    178 

trace gases can diffuse into cloud droplets;5 however, without measurement of these gases or

179 

dissolved low molecular weight species, this pathway could not be thoroughly evaluated in this

180 

work. Further reactions have been proposed to occur in evaporating cloud droplets, due to

181 

increasing solute concentrations and dehydration;29-31 however, since our chemical analysis was

182 

performed on whole water samples, these pathways were not investigated.

183 

Previous studies suggested that cloud processing would likely increase the overall

184 

oxygen/carbon (O:C) ratio of organic compounds in cloud water.5 However, all three pairs of

185 

cloud water/particle samples examined in this study showed no statistically significant difference

186 

in the average O:C ratios (0.65 ± 0.35) of the organic compounds detected in the negative mode.

187 

The O:C ratios corresponding to each SOAS sample are shown in Figure 2. Previously, Nguyen

188 

et al.32 observed a similar effect in a laboratory study of aqueous isoprene SOA dissolved in

189 

water and exposed to simulated sunlight.

190 

While the average O:C ratio of the organic material in the Nguyen et al.32 isoprene SOA

191 

study was not largely affected by the aqueous processing, photolysis of the aqueous SOA was

192 

observed to lead to oligomer fragmentation. In previous studies of isoprene oxidation and SOA

193 

formation, an inverse relationship between the number and chain length of oligomers of 2-

194 

methylglyceric acid (2-MGA) and glycolaldehyde (GLYC) and the relative humidity was

195 

observed, suggesting either fewer formation reactions within the more dilute aerosols or

196 

fragmentation due to excess water reversibly driving the likely aldol condensation reaction.33-36

197 

Nonetheless, isoprene derived oligomers have been observed previously in cloud water.6 Within

198 

our SOAS particle and cloud water samples, oligomers of 2-MGA and GLYC were identified by

199 

comparison with previous laboratory studies.33 A list of the oligomeric compounds identified for

200 

all three flights is shown in Table S3. However, while the 2-MGA and GLYC oligomers were

9   

ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 31

    201 

observed within the cloud water, they were reduced in number when compared to the particle

202 

samples. As an example, within the range of 150-400 m/z, 16 potential 2-MGA and GLYC

203 

oligomers were found in the June 27th particle sample compared to 3 observed in the cloud water.

204 

The June 10th sampling was the only case with a greater number of 2-MGA and GLYC

205 

oligomers identified in the cloud water, compared to the corresponding particle sample. In

206 

addition, the June 10th and 27th particle samples contained several trimers while the cloud water

207 

only contained dimers and monomers, based on comparison of formulas with previous studies.33

208 

Evidence of possible oligomer decomposition in the cloud water was also observed by the

209 

average number of carbons in the likely 2-MGA and GLYC oligomers across all aircraft samples

210 

decreasing from 8.1 ± 0.3 in particles to 7.4 ± 0.3 in cloud water. This observation is in

211 

agreement with the laboratory-based study of aqueous processing of isoprene SOA by Nguyen et

212 

al.32 However, rather than indicating oligomer fragmentation, it is possible that the higher m/z

213 

values in the cloud water could have been suppressed due to mass discrimination associated with

214 

the differences in the mass ranges scanned between the particle and cloud water samples.

215 

CHOS Compounds

216 

For the SOAS flights, negative ion CHOS compounds decreased in number fraction, from

217 

an average of 29% in the particles to 6% in the cloud water samples, as shown in Figure 3. This

218 

trend is consistent with the total number of CHOS compounds observed as well. On average, a

219 

factor of ~15 more unique CHOS compounds (by number) were present in the particles

220 

compared to the cloud water (Table S4). Modeling and experimental studies of the formation of

221 

organosulfates have shown that production is maximized in particles that are acidic and under

222 

low RH conditions.37, 38 Thus, as expected, no new organosulfate masses were identified within

223 

the cloud water mass spectra, compared to the corresponding particle samples. Overall, fewer

10   

ACS Paragon Plus Environment

Page 11 of 31

Environmental Science & Technology

    224 

than 8% of the CHOS particle-phase compounds were also observed in the cloud water.

225 

Therefore, this trend of decreasing diversity of CHOS compounds in the cloud water is

226 

consistent with decomposition of these compounds in the aqueous environment.

227 

Hydrolysis of organosulfates in the cloud water could account for the decreasing number

228 

and fraction of CHOS compounds, as discussed above. Previous laboratory kinetics studies have

229 

shown that tertiary organosulfate compounds hydrolyze in aqueous solution on a time scale of

230 

minutes to hours,15 which could be relevant during cloud events.39 To investigate this process for

231 

the aircraft samples, possible hydrolysis products of particle-phase organosulfates (ROSO H

232 

H O ↔ ROH

233 

difference of SO3 between the particle phase organosulfates and the compounds measured in the

234 

cloud water. Using the June 10th flight as a case study, 46 organosulfate compounds were

235 

identified in the particle sample with 30 possible corresponding hydrolysis products in the cloud

236 

water. A neutral mass spectrum of these product mass pairs is shown in Figure 4. For the particle

237 

sample, 27 of the hydrolysis product masses were also identified. The presence of these CHO

238 

compounds is not unexpected since these compounds can react to form the organosulfates also

239 

found in the particle phase.14 Similarly, for the June 13th and 27th samples, 65-72% of the CHOS

240 

compounds in the particle-phase had corresponding hydrolysis product CHO formulas in the

241 

cloud water. These organosulfates are characterized by O:C of 2.25, illustrating that they

242 

contained oxygenated functional groups beyond the sulfate group, which is expected based on

243 

previously proposed organosulfate formation pathways.40 This full analysis, for all sampling

244 

days, is included in the supplementary information (Table S5). This indicates that the hydrolysis

245 

reactions may dominate the chemistry of organosulfate compounds within cloud water and can

246 

occur on cloud lifetime timescales. For the June 10th cloud water sample, four likely

H SO

were identified by searching for compounds with a neutral mass

11   

ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 31

    247 

organosulfates corresponded to biogenic volatile organic compound oxidation products with

248 

proposed structures from previous studies.40 In previous laboratory studies, Darer et al.14

249 

reported that kinetics of the hydrolysis of tertiary organosulfates under acidic conditions were

250 

relatively fast (lifetimes of minutes to hours), while at neutral pH the reaction was much slower

251 

(hours to days). Given an expected cloud pH of ~3,41 this indicates that the presence of tertiary

252 

sulfate functionality in the cloud water samples should be rare due to hydrolysis.14 Based on this

253 

study, it is not surprising that all of the organosulfates identified in the cloud water were

254 

proposed to have primary sulfate functionalities and were not expected to hydrolyze on the

255 

timescale of a cloud event.14 Similar trends were observed for the other two SOAS flights as

256 

well. In contrast to the low abundance of tertiary organosulfates in the cloud water samples, 50-

257 

86% of the identified particle-phase organosulfates, across the three SOAS flights, were

258 

characterized by tertiary functionality with hydrolysis lifetimes of ~20-460 hours under neutral

259 

conditions.14

260 

Across all of the SOAS samples, the highest abundance organosulfate compounds

261 

observed were suggested to be isoprene oxidation products,40 consistent with the high isoprene

262 

emissions in this forested region of Alabama.42 The peak at m/z 215 (C5H11O7S1-) was

263 

ubiquitously abundant in all samples, as was also observed in a previous cloud water study above

264 

the isoprene emission region of Missouri.6 The presence of this species across all particle and

265 

cloud water samples suggests that it can be used as a tracer to confirm the influence of boundary

266 

layer particles on the organic composition of cloud droplets. This organosulfate is produced from

267 

isoprene epoxydiols (IEPOX) and contains a primary sulfate functional group,40 making it stable

268 

to hydrolysis (lifetime >2500 hours),39 which also explains the presence of this particular

269 

organosulfate in many field aerosol samples.27, 40, 43-49 Overall, of the observed organosulfates

12   

ACS Paragon Plus Environment

Page 13 of 31

Environmental Science & Technology

    270 

that could be assigned possible structures, 30-50%, by number per flight, were attributed to

271 

isoprene oxidation.40 Among the isoprene oxidation organosulfates, 2-methylglyoxal sulfate

272 

esters34 were observed in all three particle samples. However, due to the tertiary functionality of

273 

the sulfate group, these compounds were not observed in the cloud water, as expected due to

274 

their fast hydrolysis.14 In addition, many minor organosulfate compounds identified within the

275 

particles and cloud water samples were suggested to have monoterpene precursors.

276 

CHON compounds

277 

Another important difference between the particle and cloud water samples across all

278 

three sampling days was that negative ion CHON compounds significantly increased in number

279 

fraction between the particle and cloud water samples (Figure 5). For the SOAS samples, no

280 

more than five CHON compounds were observed in each of the particle samples, with ~4x more

281 

CHON compounds detected in the cloud water (Table S6). On average, the nitrogen/carbon N:C

282 

ratio of all negative ion organic compounds increased from 0.03, for the particle-phase samples,

283 

to 0.14, for the cloud water samples, with the cloud water CHON class compounds having an

284 

average N:C ratio of 0.3. The overall addition of nitrogen to the organic compounds in the cloud

285 

water could involve reactions of organic compounds with inorganic nitrogen particulate species

286 

(e.g., ammonium, nitrate, nitrite), low molecular weight nitrogen-containing particle phase

287 

organic compounds, or water-soluble nitrogen-containing gases (e.g. HNO3, N2O5, organic

288 

nitrates). While the bimolecular reactions that could be generating these CHON compounds

289 

should be slower in the dilute droplet environment, it is likely that the reaction is preceded by

290 

dissolution of a soluble trace gas that serves as reactant. A few CHONS compounds were also

291 

detected in the SOAS samples, but their appearance did not show any conclusive trends between

292 

the particle and cloud water samples.

13   

ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 31

    293 

Since most N-containing compounds preferentially form positive ions, previous studies

294 

have examined CHON compound formation in the aqueous phase using positive ion mode

295 

detection.29,

296 

assess whether the same trends observed in the negative mode were also found in the positive

297 

mode. The number of assigned compounds in the positive spectra was too small for a robust

298 

statistical analysis. However, despite this, some trends can be gleaned from the raw number of

299 

CHON compounds assigned in the positive ion mode. In two of the three SOAS flights, an

300 

increase in number of assigned CHON compounds was observed for the cloud water samples

301 

compared to the particle samples, suggesting CHON compound formation in cloud droplets. A

302 

list of assigned positive mode CHON compounds is shown in Table S2. With an average N:C

303 

ratio of 0.24, the positive mode CHON compounds detected in the cloud water display a similar

304 

N:C ratio as the negative mode CHON compounds. However, the negative mode CHON

305 

compounds detected in the SOAS cloud water samples were markedly different from the positive

306 

mode in other characteristics. Overall, the H:C ratio of the negative mode compounds was lower

307 

than the positive mode (Figure S3), indicating less saturated compounds in the negative mode. In

308 

addition, 46% of the negative mode CHON compounds were calculated to have an aromaticity

309 

index50 (AE) above 0.67, indicating that many of these compounds likely have aromatic

310 

structures. In contrast, only 8% of the positive mode compounds had AE values above 0.67,

311 

indicating many of these compounds likely do not contain aromatic structures, in line with the

312 

observation of generally saturated formulas.

30, 37

Therefore, positive mode data for the CHON compounds were examined to

The formation of CHON compounds in aqueous organic mixtures has been studied

313  314 

previously using bulk solutions in the laboratory.29,

30

De Haan et al.29 and Lee et al.30 both

315 

observed CHON compounds in aerosols following bulk solution atomization and suggested that

14   

ACS Paragon Plus Environment

Page 15 of 31

Environmental Science & Technology

    316 

evaporation increases the rates of reactions of organic compounds with amines and ammonium

317 

ions by increasingly concentrating the solution, with sulfate promoting the reactions. The

318 

observation of CHON formation within the SOAS cloud droplets suggests that droplet

319 

evaporation may not be a unique process generating N-containing compounds. In the aqueous

320 

isoprene SOA study of Nguyen et al.32 multi-nitrogen-containing compounds were observed to

321 

form following illumination of the aqueous solution, and a decrease in the number of single

322 

nitrogen-containing compounds in the aqueous solution was also observed. This observation was

323 

attributed to photolytic reaction involving organonitrates32 and could also be explained by the

324 

expected fast hydrolysis of tertiary organonitrates.14 However, for the SOAS cloud water

325 

samples, 75% of the negative mode CHON compounds and most of the positive mode CHON

326 

compounds correspond to formulas (Tables S1 and S2) with multiple nitrogen atoms and/or too

327 

few oxygen atoms to be assigned as organonitrates. Therefore, while there could be nitrate

328 

functionalities present in some of the compounds, it is more likely that the CHON species have

329 

other N-containing heteroatom moieties. The positive mode CHON compounds observed in the

330 

SOAS cloud water were characterized by an average O:C ratio of 0.44 which is within the range

331 

of bulk O:C ratios (0.3 – 0.7) measured by De Haan et al.29 Including both positive and negative

332 

mode data, these compounds are generally in line with the products observed in all three

333 

previous aqueous-phase CHON formation studies.29,

334 

formulas in the current study were found to exactly correspond to any of the compounds reported

335 

in the previous laboratory studies.29, 30, 32 Further laboratory studies are needed to elucidate the

336 

chemical mechanisms that result in the formation of these CHON compounds within cloud

337 

droplets. Previously, aromatic N-containing compounds have been found to act as chromophores

30, 32

However, none of the assigned

15   

ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 31

    338 

in brown carbon aerosol, changing the optical properties of the particles and their climate

339 

impacts.29, 32, 37

340 

Overall, the majority of the compounds observed in the SOAS particle and cloud water

341 

samples corresponded to oxidation products of isoprene and monoterpenes, as anticipated due to

342 

the significant biogenic volatile organic compound emissions from the local forested region.

343 

While the average properties (e.g. O:C ratios) of high molecular weight compounds in the

344 

measured particles and cloud water did not differ significantly, aqueous-phase cloud processing

345 

was observed to lead to organosulfate hydrolysis and CHON compound formation. This

346 

indicates that cloud processing of organic aerosol could play a significant role in changing the

347 

chemical properties of the aerosol following cloud evaporation. The aqueous-phase formation of

348 

nitrogen- containing organic compounds, which may have light absorbing properties,30, 32, 34, 37

349 

could affect the radiative forcing budget in environments influenced by both biogenic and

350 

anthropogenic emissions, and future laboratory studies should focus on identification of the

351 

mechanisms responsible for this type of chemistry. Future modeling studies which add reaction

352 

pathways which hydrolyze tertiary organosulfates and generate nitrogen-containing organics

353 

through aqueous processing should improve more accurate estimations of cloud and aerosol

354 

chemistry.

355  356 

Acknowledgments

357 

Funding for SOAS sampling was provided by NSF (AGS-1228496) and EPA

358 

(R835409). High-resolution mass spectrometry analyses were performed at the Environmental

359 

Molecular Sciences Laboratory (EMSL), a national scientific user facility located at the Pacific

360 

Northwest National Laboratory (PNNL) and sponsored by the Office of Biological and

16   

ACS Paragon Plus Environment

Page 17 of 31

Environmental Science & Technology

    361 

Environmental Research of the U.S Department of Energy (DOE). PNNL is operated for DOE

362 

by Battelle Memorial Institute under Contract No. DE-AC06-76RL0 1830. Travel funds to

363 

PNNL were provided by the University of Michigan College of Literature, Science, and the Arts

364 

and Department of Chemistry.

365  366 

Supporting Information Negative ion mass spectra and tables of assigned compounds, including identified

367  368 

oligomers and organosulfates, are provided. This material is available free of charge via the

369 

Internet at http://pubs.acs.org.

17   

ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 31

    370 

References

371 

1.

effects. Angew. Chem. Int. Ed. Engl. 2005, 44, 7520-40, doi:10.1002/anie.200501122.

372  373 

2.

Andreae, M.; Rosenfeld, D., Aerosol–cloud–precipitation interactions. Part 1. The nature and sources of cloud-active aerosols. Earth-Sci. Rev. 2008, 89, 13-41.

374  375 

Poschl, U., Atmospheric aerosols: composition, transformation, climate and health

3.

Hallquist, M.; Wenger, J. C.; Baltensperger, U.; Rudich, Y.; Simpson, D.; Claeys, M.;

376 

Dommen, J.; Donahue, N. M.; George, C.; Goldstein, A. H.; Hamilton, J. F.; Herrmann,

377 

H.; Hoffmann, T.; Iinuma, Y.; Jang, M.; Jenkin, M. E.; Jimenez, J. L.; Kiendler-Scharr,

378 

A.; Maenhaut, W.; McFiggans, G.; Mentel, T. F.; Monod, A.; Prevot, A. S. H.; Seinfeld,

379 

J. H.; Surratt, J. D.; Szmigielski, R.; Wildt, J., The formation, properties and impact of

380 

secondary organic aerosol: current and emerging issues. Atmos. Chem. Phys. 2009, 9,

381 

5155-5236.

382 

4.

Carlton, A. G.; Turpin, B. J.; Altieri, K. E.; Seitzinger, S. P.; Mathur, R.; Roselle, S. J.;

383 

Weber, R. J., CMAQ Model Performance Enhanced When In-Cloud Secondary Organic

384 

Aerosol is Included: Comparisons of Organic Carbon Predictions with Measurements.

385 

Environ. Sci. Technol. 2008, 42, 8798-8802, doi:10.1021/Es801192n.

386 

5.

Ervens, B.; Turpin, B. J.; Weber, R. J., Secondary organic aerosol formation in cloud

387 

droplets and aqueous particles (aqSOA): a review of laboratory, field and model studies.

388 

Atmos. Chem. Phys. 2011, 11, 11069-11102, doi:10.5194/acp-11-11069-2011.

389 

6.

Pratt, K. A.; Fiddler, M. N.; Shepson, P. B.; Carlton, A. G.; Surratt, J. D., Organosulfates

390 

in cloud water above the Ozarks’ isoprene source region. Atmos. Environ. 2013, 77, 231-

391 

238.

18   

ACS Paragon Plus Environment

Page 19 of 31

Environmental Science & Technology

    392 

7.

Feng, J. S.; Moller, D., Characterization of water-soluble macromolecular substances in

393 

cloud

394 

doi:10.1023/B:Joch.0000044377.93748.E4.

395 

8.

water.

J.

Atmos.

Chem.

2004,

217-233,

48,

Lee, A. K. Y.; Hayden, K. L.; Herckes, P.; Leaitch, W. R.; Liggio, J.; Macdonald, A. M.;

396 

Abbatt, J. P. D., Characterization of aerosol and cloud water at a mountain site during

397 

WACS 2010: secondary organic aerosol formation through oxidative cloud processing.

398 

Atmos. Chem. Phys. 2012, 12, 7103-7116, doi:10.5194/acp-12-7103-2012.

399 

9.

Samy, S.; Mazzoleni, L. R.; Mishra, S.; Zielinska, B.; Hallar, A. G., Water-soluble

400 

organic compounds at a mountain-top site in Colorado, USA. Atmos. Environ. 2010, 44,

401 

1663-1671, doi:10.1016/j.atmosenv.2010.01.033.

402 

10.

van Pinxteren, D.; Herrmann, H., Determination of functionalised carboxylic acids in

403 

atmospheric particles and cloud water using capillary electrophoresis/mass spectrometry.

404 

J. Chromatogr. A 2007, 1171, 112-123.

405 

11.

Lee, A. K. Y.; Herckes, P.; Leaitch, W. R.; Macdonald, A. M.; Abbatt, J. P. D., Aqueous

406 

OH oxidation of ambient organic aerosol and cloud water organics: Formation of highly

407 

oxidized

408 

doi:10.1029/2011gl047439.

409 

12.

products.

Geophysical

Research

Letters

2011,

38,

L11805,

McNeill, V. F., Aqueous organic chemistry in the atmosphere: sources and chemical

410 

processing of organic aerosols. Environ. Sci. Technol. 2015, 49, 1237-44,

411 

doi:10.1021/es5043707.

412 

13.

Hu, K.; Darer, A. I.; Elrod, M. J., Thermodynamics and kinetics of the hydrolysis of

413 

atmospherically relevant organonitrates and organosulfates. Atmos. Chem. Phys. 2011,

414 

11, 8307-8320.

19   

ACS Paragon Plus Environment

Environmental Science & Technology

Page 20 of 31

    415 

14.

Darer, A. I.; Cole-Filipiak, N. C.; O'Connor, A. E.; Elrod, M. J., Formation and stability

416 

of atmospherically relevant isoprene-derived organosulfates and organonitrates. Environ.

417 

Sci. Technol. 2011, 45, 1895-902, doi:10.1021/es103797z.

418 

15.

Goldstein, A. H.; Koven, C. D.; Heald, C. L.; Fung, I. Y., Biogenic carbon and

419 

anthropogenic pollutants combine to form a cooling haze over the southeastern United

420 

States. Proc Natl Acad Sci U S A 2009, 106, 8835-40, doi:10.1073/pnas.0904128106.

421 

16.

Weber, R. J.; Sullivan, A. P.; Peltier, R. E.; Russell, A.; Yan, B.; Zheng, M.; De Gouw,

422 

J.; Warneke, C.; Brock, C.; Holloway, J. S., A study of secondary organic aerosol

423 

formation in the anthropogenic‐influenced southeastern United States. J. Geophys. Res.

424 

Atmos. 2007, 112, D13302.

425 

17.

Carlton, A. G.; Turpin, B. J., Particle partitioning potential of organic compounds is

426 

highest in the Eastern US and driven by anthropogenic water. Atmos. Chem. Phys. 2013,

427 

13, 10203-10214, doi:10.5194/acp-13-10203-2013.

428 

18.

Bateman, A. P.; Nizkorodov, S. A.; Laskin, J.; Laskin, A., Time-resolved molecular

429 

characterization of limonene/ozone aerosol using high-resolution electrospray ionization

430 

mass spectrometry. PCCP 2009, 11, 7931-7942, doi:10.1039/B905288g.

431 

19.

Huebert, B. J.; Vanbramer, S.; Tschudy, K. L., Liquid Cloudwater Collection Using

432 

Modified

433 

doi:10.1007/Bf00053859.

434 

20.

Mohnen

Slotted

Rods.

J.

Atmos.

Chem.

1988,

6,

251-263,

Hill, K. A.; Shepson, P. B.; Galbavy, E. S.; Anastasio, C.; Kourtev, P. S.; Konopka, A.;

435 

Stirm, B. H., Processing of atmospheric nitrogen by clouds above a forest environment. J.

436 

Geophys. Res. Atmos. 2007, 112, D11301, doi:10.1029/2006jd008002.

20   

ACS Paragon Plus Environment

Page 21 of 31

Environmental Science & Technology

    437 

21.

Cloud-Water Collector in Summer Clouds. J. Atmos. Oceanic Tech. 1992, 9, 35-41.

438  439 

Kim, Y. J.; Boatman, J. F., The Collection Efficiency of a Modified Mohnen Slotted-Rod

22.

Roach, P. J.; Laskin, J.; Laskin, A., Nanospray desorption electrospray ionization: an

440 

ambient method for liquid-extraction surface sampling in mass spectrometry. Analyst

441 

2010, 135, 2233-6, doi:10.1039/c0an00312c.

442 

23.

Roach, P. J.; Laskin, J.; Laskin, A., Molecular Characterization of Organic Aerosols

443 

Using Nanospray-Desorption/Electrospray Ionization-Mass Spectrometry†. Anal. Chem.

444 

2010, 82, 7979-7986.

445 

24.

complex organic mixtures. Anal. Chem. 2011, 83, 4924-9, doi:10.1021/ac200654j.

446  447 

Roach, P. J.; Laskin, J.; Laskin, A., Higher-order mass defect analysis for mass spectra of

25.

Bateman, A. P.; Laskin, J.; Laskin, A.; Nizkorodov, S. A., Applications of high-

448 

resolution electrospray ionization mass spectrometry to measurements of average oxygen

449 

to carbon ratios in secondary organic aerosols. Environ. Sci. Technol. 2012, 46, 8315-24,

450 

doi:10.1021/es3017254.

451 

26.

O’Brien, R. E.; Laskin, A.; Laskin, J.; Liu, S.; Weber, R.; Russell, L. M.; Goldstein, A.

452 

H.,

453 

desorption/electrospray ionization mass spectrometry: CalNex 2010 field study. Atmos.

454 

Environ. 2013, 68, 265-272, doi:10.1016/j.atmosenv.2012.11.056.

455 

27.

Molecular

characterization

of

organic

aerosol

using

nanospray

Tao, S.; Lu, X.; Levac, N.; Bateman, A. P.; Nguyen, T. B.; Bones, D. L.; Nizkorodov, S.

456 

A.; Laskin, J.; Laskin, A.; Yang, X., Molecular characterization of organosulfates in

457 

organic aerosols from Shanghai and Los Angeles urban areas by nanospray-desorption

458 

electrospray ionization high-resolution mass spectrometry. Environ. Sci. Technol. 2014,

459 

48, 10993-1001, doi:10.1021/es5024674.

21   

ACS Paragon Plus Environment

Environmental Science & Technology

Page 22 of 31

    460 

28.

Zhao, Y.; Hallar, A. G.; Mazzoleni, L. R., Atmospheric organic matter in clouds: exact

461 

masses and molecular formula identification using ultrahigh-resolution FT-ICR mass

462 

spectrometry. Atmos. Chem. Phys. 2013, 13, 12343-12362, doi:10.5194/acp-13-12343-

463 

2013.

464 

29.

De Haan, D. O.; Hawkins, L. N.; Kononenko, J. A.; Turley, J. J.; Corrigan, A. L.;

465 

Tolbert, M. A.; Jimenez, J. L., Formation of nitrogen-containing oligomers by

466 

methylglyoxal and amines in simulated evaporating cloud droplets. Environ. Sci.

467 

Technol. 2011, 45, 984-91, doi:10.1021/es102933x.

468 

30.

Lee, A. K.; Zhao, R.; Li, R.; Liggio, J.; Li, S. M.; Abbatt, J. P., Formation of light

469 

absorbing organo-nitrogen species from evaporation of droplets containing glyoxal and

470 

ammonium sulfate. Environ. Sci. Technol. 2013, 47, 12819-26, doi:10.1021/es402687w.

471 

31.

Loeffler, K. W.; Koehler, C. A.; Paul, N. M.; De Haan, D. O., Oligomer formation in

472 

evaporating aqueous glyoxal and methyl glyoxal solutions. Environ. Sci. Technol. 2006,

473 

40, 6318-6323.

474 

32.

Nguyen, T. B.; Laskin, A.; Laskin, J.; Nizkorodov, S. A., Direct aqueous photochemistry

475 

of isoprene high-NO(x) secondary organic aerosol. Phys. Chem. Chem. Phys. 2012, 14,

476 

9702-14, doi:10.1039/c2cp40944e.

477 

33.

Nguyen, T. B.; Roach, P. J.; Laskin, J.; Laskin, A.; Nizkorodov, S. A., Effect of humidity

478 

on the composition of isoprene photooxidation secondary organic aerosol. Atmos. Chem.

479 

Phys. 2011, 11, 6931-6944, doi:10.5194/acp-11-6931-2011.

480 

34.

Birdsall, A. W.; Zentner, C. A.; Elrod, M. J., Study of the kinetics and equilibria of the

481 

oligomerization reactions of 2-methylglyceric acid. Atmos. Chem. Phys. 2013, 13, 3097-

482 

3109, doi:10.5194/acp-13-3097-2013.

22   

ACS Paragon Plus Environment

Page 23 of 31

Environmental Science & Technology

    483 

35.

Zhang, H.; Surratt, J. D.; Lin, Y. H.; Bapat, J.; Kamens, R. M., Effect of relative humidity

484 

on SOA formation from isoprene/NO photooxidation: enhancement of 2-methylglyceric

485 

acid and its corresponding oligoesters under dry conditions. Atmos. Chem. Phys. 2011,

486 

11, 6411-6424, doi:10.5194/acp-11-6411-2011.

487 

36.

Liu, Y.; Siekmann, F.; Renard, P.; El Zein, A.; Salque, G.; El Haddad, I.; Temime-

488 

Roussel, B.; Voisin, D.; Thissen, R.; Monod, A., Oligomer and SOA formation through

489 

aqueous phase photooxidation of methacrolein and methyl vinyl ketone. Atmos. Environ.

490 

2012, 49, 123-129, doi:10.1016/j.atmosenv.2011.12.012.

491 

37.

Nguyen, T. B.; Lee, P. B.; Updyke, K. M.; Bones, D. L.; Laskin, J.; Laskin, A.;

492 

Nizkorodov, S. A., Formation of nitrogen‐and sulfur‐containing light‐absorbing

493 

compounds accelerated by evaporation of water from secondary organic aerosols. J.

494 

Geophys. Res. Atmos. 2012, 117, D01207, doi:10.1029/2011JD016944.

495 

38.

McNeill, V. F.; Woo, J. L.; Kim, D. D.; Schwier, A. N.; Wannell, N. J.; Sumner, A. J.;

496 

Barakat, J. M., Aqueous-phase secondary organic aerosol and organosulfate formation in

497 

atmospheric aerosols: a modeling study. Environ. Sci. Technol. 2012, 46, 8075-81,

498 

doi:10.1021/es3002986.

499 

39.

Darer, A. I.; Cole-Filipiak, N. C.; O’Connor, A. E.; Elrod, M. J., Formation and stability

500 

of atmospherically relevant isoprene-derived organosulfates and organonitrates.

501 

Environmental science & technology 2011, 45, 1895-1902.

502 

40.

Surratt, J. D.; Gomez-Gonzalez, Y.; Chan, A. W.; Vermeylen, R.; Shahgholi, M.;

503 

Kleindienst, T. E.; Edney, E. O.; Offenberg, J. H.; Lewandowski, M.; Jaoui, M.;

504 

Maenhaut, W.; Claeys, M.; Flagan, R. C.; Seinfeld, J. H., Organosulfate formation in

23   

ACS Paragon Plus Environment

Environmental Science & Technology

Page 24 of 31

    505 

biogenic secondary organic aerosol. J. Phys. Chem. A 2008, 112, 8345-78,

506 

doi:10.1021/jp802310p.

507 

41.

Collett Jr, J. L.; Bator, A.; Sherman, D. E.; Moore, K. F.; Hoag, K. J.; Demoz, B. B.; Rao,

508 

X.; Reilly, J. E., The chemical composition of fogs and intercepted clouds in the United

509 

States. Atmospheric Research 2002, 64, 29-40, doi:http://dx.doi.org/10.1016/S0169-

510 

8095(02)00077-7.

511 

42.

Guenther, A.; Karl, T.; Harley, P.; Wiedinmyer, C.; Palmer, P. I.; Geron, C., Estimates of

512 

global terrestrial isoprene emissions using MEGAN (Model of Emissions of Gases and

513 

Aerosols from Nature). Atmos. Chem. Phys. 2006, 6, 3181-3210.

514 

43.

Chan, M. N.; Surratt, J. D.; Claeys, M.; Edgerton, E. S.; Tanner, R. L.; Shaw, S. L.;

515 

Zheng, M.; Knipping, E. M.; Eddingsaas, N. C.; Wennberg, P. O.; Seinfeld, J. H.,

516 

Characterization and quantification of isoprene-derived epoxydiols in ambient aerosol in

517 

the

518 

doi:10.1021/es100596b.

519 

44.

southeastern

United

States.

Environ.

Sci.

Technol.

2010,

44,

4590-6,

Froyd, K. D.; Murphy, S. M.; Murphy, D. M.; de Gouw, J. A.; Eddingsaas, N. C.;

520 

Wennberg, P. O., Contribution of isoprene-derived organosulfates to free tropospheric

521 

aerosol

522 

doi:10.1073/pnas.1012561107.

523 

45.

mass.

Proc

Natl

Acad

Sci

U

S

A

2010,

107,

21360-5,

Gómez‐González, Y.; Surratt, J. D.; Cuyckens, F.; Szmigielski, R.; Vermeylen, R.; Jaoui,

524 

M.; Lewandowski, M.; Offenberg, J. H.; Kleindienst, T. E.; Edney, E. O.,

525 

Characterization of organosulfates from the photooxidation of isoprene and unsaturated

526 

fatty acids in ambient aerosol using liquid chromatography/(−) electrospray ionization

527 

mass spectrometry. J. Mass Spectrom. 2008, 43, 371-382.

24   

ACS Paragon Plus Environment

Page 25 of 31

Environmental Science & Technology

    528 

46.

Hatch, L. E.; Creamean, J. M.; Ault, A. P.; Surratt, J. D.; Chan, M. N.; Seinfeld, J. H.;

529 

Edgerton, E. S.; Su, Y. X.; Prather, K. A., Measurements of Isoprene-Derived

530 

Organosulfates in Ambient Aerosols by Aerosol Time-of-Flight Mass Spectrometry - Part

531 

1: Single Particle Atmospheric Observations in Atlanta. Environ. Sci. Technol. 2011, 45,

532 

5105-5111, doi:10.1021/Es103944a.

533 

47.

Yttri, K. E.; Simpson, D.; Nojgaard, J. K.; Kristensen, K.; Genberg, J.; Stenstrom, K.;

534 

Swietlicki, E.; Hillamo, R.; Aurela, M.; Bauer, H.; Offenberg, J. H.; Jaoui, M.; Dye, C.;

535 

Eckhardt, S.; Burkhart, J. F.; Stohl, A.; Glasius, M., Source apportionment of the summer

536 

time carbonaceous aerosol at Nordic rural background sites. Atmos. Chem. Phys. 2011,

537 

11, 13339-13357, doi:10.5194/acp-11-13339-2011.

538 

48.

Stone, E. A.; Yang, L. M.; Yu, L. Y. E.; Rupakheti, M., Characterization of

539 

organosulfates in atmospheric aerosols at Four Asian locations. Atmos. Environ. 2012,

540 

47, 323-329, doi:10.1016/j.atmosenv.2011.10.058.

541 

49.

O'Brien, R. E.; Laskin, A.; Laskin, J.; Rubitschun, C. L.; Surratt, J. D.; Goldstein, A. H.,

542 

Molecular Characterization of S‐and N‐containing Organic Constituents in Ambient

543 

Aerosols by Negative Ion Mode High‐Resolution Nanospray Desorption Electrospray

544 

Ionization Mass Spectrometry: CalNex 2010 field study. J. Geophys. Res. Atmos. 2014,

545 

119, 12,706-12,720, doi:10.1002/2014JD021955.

546 

50.

Koch, B. P.; Dittmar, T., From mass to structure: an aromaticity index for high-resolution

547 

mass data of natural organic matter. Rapid Commun. Mass Spectrom. 2006, 20, 926-932,

548 

doi:10.1002/Rcm.2386.

549 

25   

ACS Paragon Plus Environment

Environmental Science & Technology

Page 26 of 31

   

550  551 

Figure 1: Negative ion mode neutral mass spectra for (a) particle and (b) cloud water samples

552 

collected on June 10th. Assignment class is indicated by the color of the bars. Inset pie charts

553 

indicate the number fraction of each class assigned for the samples.

554 

26   

ACS Paragon Plus Environment

Page 27 of 31

Environmental Science & Technology

   

555  556 

Figure 2: Average O:C ratios characteristic for organic compounds detected in negative ion

557 

mode mass spectra of particle and cloud water samples. Error bars represent the standard error.

558 

The total number of assigned compounds in the spectra is displayed above the bars.

559 

27   

ACS Paragon Plus Environment

Environmental Science & Technology

Page 28 of 31

   

560  561 

Figure 3: Number fractions of CHOS compounds, detected in the negative ion mode, present in

562 

the SOAS particle and cloud water samples. Error bars represent the standard error of the

563 

fraction. The total number of assigned CHOS compounds in the negative spectra is displayed

564 

above the bars.

565 

28   

ACS Paragon Plus Environment

Page 29 of 31

Environmental Science & Technology

   

566  567 

Figure 4: Neutral mass spectrum of particle-phase CHOS compounds and potential

568 

corresponding hydrolysis products, in the June 10 samples. Assignments of intense peaks are

569 

noted.

29   

ACS Paragon Plus Environment

Environmental Science & Technology

Page 30 of 31

   

570  571 

Figure 5: Number fractions of CHON compounds, detected in the negative ion mode, for the

572 

SOAS particle and cloud water samples. Total CHON fraction error bars represent the standard

573 

error of the fraction. The total number of assigned CHON compounds in the negative spectra is

574 

displayed above the bars.

575 

30   

ACS Paragon Plus Environment

Page 31 of 31

Environmental Science & Technology

   

576  577 

TOC Cover Art

31   

ACS Paragon Plus Environment