Synthesis of Four Monoterpene-Derived Organosulfates and Their

†Environmental Science Programs, §Institute of Environment, and ‡Department of Chemistry, Hong Kong University of Science & Technology, Hong Kong...
1 downloads 12 Views 803KB Size
Subscriber access provided by CORNELL UNIVERSITY LIBRARY

Article

Synthesis of Four Monoterpene-derived Organosulfates and their Quantification in Atmospheric Aerosol Samples Yuchen Wang, Jingyun Ren, X. H. Hilda Huang, Rongbiao Tong, and Jian Zhen Yu Environ. Sci. Technol., Just Accepted Manuscript • Publication Date (Web): 26 May 2017 Downloaded from http://pubs.acs.org on May 27, 2017

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 28

Environmental Science & Technology

1  2 

Synthesis of Four Monoterpene-derived Organosulfates and their Quantification in Atmospheric Aerosol Samples

3  4  5 

Yuchen Wang,† Jingyun Ren,‡ X. H. Hilda Huang,§ Rongbiao Tong,*,‡ Jian Zhen Yu*,†,‡ † Environmental Science Programs, Hong Kong University of Science & Technology, Hong Kong, China

6  7 

§

Institute of Environment, Hong Kong University of Science & Technology, Hong Kong,

China

8  9  10  11  12  13 

‡ Department of Chemistry, Hong Kong University of Science & Technology, Hong Kong, China *Corresponding Authors Phone: (852) 2358 7389, Fax: (852) 2358 1594, Email: [email protected], [email protected]   Abstract

14 

Monoterpenes, a major class of biogenic volatile organic compounds, are known to produce

15 

oxidation products that further react with sulfate to form organosulfates. The accurate

16 

quantification of monoterpene-derived organosulfates (OSs) is necessary for quantifying this

17 

controllable aerosol source; however, it has been hampered by a lack of authentic standards.

18 

Here we report a unified synthesis strategy starting from the respective monoterpene through

19 

Upjohn dihydroxylation or Sharpless asymmetric dihydroxylation followed by mono-sulfation

20 

with sulfur trioxide-pyridine complex. We demonstrate the successful synthesis of four

21 

monoterpene-derived OS compounds, including -pinene OS, β-pinene OS, limonene OS and

22 

limonaketone OS. Quantification of OSs is commonly achieved using liquid chromatography-

23 

mass spectrometry (LC-MS) by either monitoring the [M-H]- ion or through multiple reaction

24 

monitoring (MRM) of mass transitions between the [M-H]- and m/z 97 ions. Comparison

25 

between the synthesized standards and previously adopted quantification surrogates reveals

26 

that camphor-10-sulfonic acid is a better quantification surrogate using [M-H]- as the

27 

quantification ion while the highly compound-specific nature of MRM quantification makes it

28 

difficult to choose a suitable surrogate. Both could be rationalized in accordance to their

29 

respective MS quantification mechanisms. The in-house availability of the authentic standards

30 

enables us to discover that β-pinene OS, due to the sulfate group at the primary carbon, partially

31 

degrades to a dehydrogenated OS compound during LC/MS analysis and a hydroperoxy OS

32 

over a prolonged storage period (> 5 month), and forms a regioisomer through intermolecular

33 

isomerization. Limonene OS was positively identified for the first time in ambient samples and

34 

found to be more abundant than -/-pinene OS in the Pearl River Delta, China. This work



 

ACS Paragon Plus Environment

Environmental Science & Technology

35 

highlights the critical importance of having authentic standards in advancing our understanding

36 

of the interactions between biogenic VOC emissions and anthropogenic sulfur pollution.

37  38 

TOC graph

39  40  41 

1. Introduction

42 

Biogenic volatile organic compounds (BVOCs), such as isoprene, monoterpenes,

43 

sesquiterpenes, contribute significantly to the global secondary organic aerosol (SOA) budget1,

44 

2

45 

promoted in the presence of high anthropogenic emissions3-5. One such a pathway is the

46 

interaction of BVOC oxidation products with sulfate particles leading to formation of

47 

organosulfates (OSs) compounds. Such OSs derived from BVOC precursors have been

48 

observed in both chamber-generated6-14 and field-collected aerosol samples15-20. These BVOC-

49 

derived OSs represent controllable biogenic SOA 21, as their formation would be diminished

50 

with the reduction in sulfate. Hence, this portion of SOA mass needs to be considered when

51 

assessing effectiveness of SO2 reduction measures on particulate matter (PM) pollution level.

. Recent research has shown that the conversion of BVOCs to SOA can be significantly

52 

Quantification of key biogenic OSs is a necessary step in assessing their contribution to the

53 

effects of PM on climate and public health and the extent of controllable biogenic SOA; however,

54 

a lack of commercially available OS standards presents a significant obstacle. The OS

55 

concentrations reported in the majority of studies in the literature are only estimates obtained

56 

through using one or more surrogate compounds that share chromatographic and mass

57 

spectrometric behaviors as close as possible to the target analytes (Table S1) in the liquid

58 

chromatography-mass spectrometric (LC/MS) analysis under negative electrospray ionization

59 

(ESI-) mode. A few alkyl OSs, including octyl sulfate 22-24, ethylhexyl sulfate 25, and camphor

60 

sulfonic acid 13, 26-28 are among the commonly adopted surrogates in quantification of BVOC-

61 

derived OSs (Table S1). The LC/MS instrument response factors of two structurally-similar

62 

aromatic OSs differed by a factor of 4.337 while the response factor of 2-methyltetrol sulfate 2 

 

ACS Paragon Plus Environment

Page 2 of 28

Page 3 of 28

Environmental Science & Technology

63 

esters34 is a factor of 2.25 lower than that of sodium propyl sulfate used as a surrogate standard18.

64 

It remains unknown how close the less polar alkyl sulfate surrogates represent the LC/MS

65 

response by the BVOC-derived OSs or whether the sulfonic acid functional group (-SO3)

66 

imparts similar ESI efficiency to the sulfate group (-OSO3-). As such, the degree of uncertainty

67 

for the concentration estimates of the BVOC-derived OSs is unknown. It is clear that lack of

68 

authentic standards has become a bottleneck in achieving accurate quantitation of individual

69 

OSs.

70 

There have been a small number of synthetic efforts, creating authentic quantification

71 

standards of OSs derived from isoprene oxidation products of first and later generations, such

72 

as glycolic acid sulfate29-32, lactic acid sulfate29,30,32,33, hydroxyactone sulfate30,31,33, 2-

73 

methyltetrol sulfate32-34, 2-methylglyceric acid sulfate32. Two studies reported synthesis of two

74 

monoterpene-derived OSs, β-pinene OS35 and limonene OS36, however, their purity information

75 

was not available from the papers.

76 

In this study, we describe a unified approach for synthesis of monoterpene-derived OSs

77 

of high purity (Fig. 1), which enables us to discover one of the synthesized OSs partially

78 

degrades to four OSs, underlining the importance of having available authentic standards. By

79 

comparing the LC/MS responses of authentic OS standards and those of the common surrogates,

80 

we provide some guidelines for selecting surrogates in LC/MS analysis for OS compounds

81 

when standards are not commercially available.

82  83 

2. Experimental

84 

All commercial compounds used in synthesis and analyses are of higher than 95% purity.

85 

The list of chemicals and their sources are provided in Sections 1 and 2 in the Supplementary

86 

Information (SI). The operation conditions of instruments (NMR and LC/MS) used in this work

87 

are also included in SI.

88 

Synthesis. A unified strategy (Fig. 2) was developed to prepare three monoterpene-

89 

derived organosulfates from the respective monoterpenes (i.e., α-pinene, β-pinene, and

90 

limonene). Specifically, monoterpenes underwent diastereoselective dihydroxylation to

91 

produce the respective diols, which were then treated with the commercially available sulfur

92 

trioxide-pyridine complex (SO3·py) to afford the O-monosulfate products 41,42 (3a, 3b, 3c and

93 

3c’ in Fig. 2) in 51-94% yield. Two dihydroxylation methods, Upjohn dihydroxylation and

94 

Sharpless asymmetric dihydroxylation, were explored (Section 4 in SI). In Upjohn

95 

dihydroxylation, a catalytic amount of osmium tetraoxide38 and stoichiometric N-methyl

96 

morphine N-oxide (NMO) as the terminal oxidant were mixed with monoterpenes to provide 3 

 

ACS Paragon Plus Environment

Environmental Science & Technology

Page 4 of 28

97 

the corresponding diols. The steric hindrance from the bicyclic structure of /-pinene led to

98 

formation of the diol as a single diastereomer (i.e., 2a and 2b, respectively, in Fig. 2), while the

99 

pseudo-equatorial isopropenyl group of limonene has a poor stereocontrol on Upjohn

100 

dihydroxylation and led to formation of a mixture of two diastereomeric diols (2c and 2c’, dr

101 

2:1)39. Sharpless asymmetric dihydroxylation40 was then employed for limonene using AD

102 

mix-β to provide the diol 2c with excellent diastereo- and enantioselectivity (dr 20:1).

103 

Oxidative cleavage of the double bond of O-monosulfates 3c and 3c’ by ozonolysis

104 

provided new O-monosulfates 3d and 3d’ with 73% yield. All the synthesized standards are

105 

stored in clear round flask without special shading from light at 4 oC in a freezer.

43,44

Structural characterization of the synthesized diols and OSs. The intermediate diol

106  107 

products and the OSs were characterized by 1H NMR (400 MHz),

108 

Distortionless Enhancement by Polarization Transfer (DEPT)-135, and heteronuclear single

109 

quantum coherence spectroscopy (HSQC). The NMR spectra are included in Figs. S3-S9 and

110 

the chemical shift values are listed in Section 8 in SI. The introduction of the sulfate group

111 

results in a downfield shift in both 1H- and 13C-NMR due to the deshielding effect of sulfate.

112 

For example, proton resonance at 3.46 ppm of -pinene diol is shifted to 3.94 ppm in 1H-NMR

113 

of -pinene OS (Fig. S7a) and the corresponding carbon signal at 70.6 ppm is found at 75.7

114 

ppm in

115 

negative carbon signal at 75.7 ppm suggests it is derived from a methylene carbon. This NMR

116 

analysis concludes that the sulfation reaction of the -pinene diol occur regioselectively at the

117 

primary alcohol. Meanwhile, the HSQC (H-C one bond correlation) spectral analysis (Fig. S7d)

118 

shows the protons resonating at 3.94 ppm is bonded to the carbon of 75.7 ppm, which reinforces

119 

our assignments of the NMR signals to the respective sulfated products. The other

120 

organosulfates derived from-pinene, limonene, and limonaketone exhibits similar chemical

121 

shift changes in 1H-NMR and 13C-NMR as -pinene OS. For example, we proposed similar

122 

deshielding effects of the sulfate on chemical shifts of the limonene OS derived from Sharpless

123 

asymmetric dihydroxylation and sulfation. The proton resonance at 3.36 ppm of limonene diol

124 

(2c) is shifted to 4.11 ppm upon sulfation (3c) (Fig. S8a), and the corresponding carbon signal

125 

at 75.1 ppm (2c) is found after sulfation at 83.9 ppm in 3c (Fig. S8b). The minor isomer 3c’

126 

was proposed to be arisen from the minor isomer in the Sharpless asymmetric dihydroxylation.

127 

The molecular masses of the O-monosulfates were confirmed with MALDI-MS analysis by

128 

directly infusing the product solutions (dissolved in methanol) into the mass spectrometer.

129 

Mass scans were performed in positive ionization mode using 2,5-dihydroxybenzoic acid as

13

C-NMR (100 MHz),

C-NMR (Fig. S7b). In DEPT-135 of the synthesized -pinene OS (Fig. S7c), the



 

13

ACS Paragon Plus Environment

Page 5 of 28

130 

Environmental Science & Technology

the matrix and the results are shown in Table 1.

131 

Determination of purity. The purity of all the synthesized OSs were determined to be

132 

higher than 99% with NMR spectra. -pinene OS, limonene OS and limonaketone OS did not

133 

show any degradation/decomposition when stored in solid form over two years; however,

134 

degradation of -pinene OS was observed. We determined the purity of -pinene OS using 1H

135 

NMR analysis of a mixture of the degraded -pinene OS sample (41.2 mg) with 15.3 mg

136 

dichloroacetic acid in CD3OD solvent as an internal standard. The relative integral values of

137 

the characteristic chemical shift peaks were used to determine the purity of -pinene OS, and

138 

we observed 19% of degradation product at five months after -pinene OS was synthesized

139 

(Section 3 in SI).

140 

LC/MS analysis. Two types of LC/MS systems were used for OSs quantification. One is

141 

a Dionex Ultimate 3000 HPLC system coupled to an ion-trap mass spectrometer (amaZon SL,

142 

Bruker Daltonics). The ESI source was operated in negative ionization mode under the

143 

following conditions: nebulizer pressure at 25 psi, dry gas flow at 10.0 L min-1, source voltage

144 

at 3.5 kV and temperature at 320 oC. The m/z range was 70 to 600 Da. LC separation was

145 

carried out on an Acquity UPLC HSS T3 column (2.1 mm×100 mm, 1.8 µm particle size;

146 

Waters, Milford, MA, USA) with a VanGuard column (HSS T3, 1.8 µm) at a flow rate of 0.3

147 

mL/min. The mobile phase consisted of water (eluent A) and methanol (eluent B), each

148 

containing 0.1% acetic acid. The gradient elution program was as follows: eluent B initially

149 

was set at 1% for 2.7 min, increased to 54% in 15.2 min and held for 1 min, then increased to

150 

90% in 7.5 min and held for 0.2 min, finally decreased to 1% in 1.8 min and held for 9.3 min.

151 

The injection volume was 5 µL. Calibration curves on this LC/MS were based on the [M-H]-

152 

molecular ions in the extracted ion chromatogram (EIC) (abbreviated as EIC quantification

153 

thereafter) and the concentrations ranged from 10 to 1000 ng/ mL (average R2>0.997). D17-

154 

octyl sulfate was used as internal standard.

155 

The second type of LC/MS system is Agilent LC-QTRAP mass spectrometers with a

156 

TurboIonSpray ionization source operated in ESI- mode. Two such LC/MS systems were used,

157 

one equipped with a QTRAP4000 mass spectrometer and the second equipped with a

158 

QTRAP4500 mass spectrometer. The MS was operated in multiple-reaction monitoring (MRM)

159 

mode for OS quantification. The MS parameters for MRM quantification were individually

160 

optimized for the mass transition between the [M-H]- ion and m/z 97 fragment ion for each OS

161 

or m/z 80 fragment ion for camphor sulfonic acid (Table S3). The dwell time for each transition

162 

was set to 100 ms. The optimal MS parameters for the three monoterpene OSs were similar



 

ACS Paragon Plus Environment

Environmental Science & Technology

163 

and the set of parameters obtained with -pinene OS was later adopted for the MRM

164 

quantification of -pinene and limonene OSs. Chromatographic separation was performed on

165 

an Acquity UPLC HSS T3 column (2.1 mm×100 mm, 1.8 µm particle size; Waters, Milford,

166 

MA, USA) with Van Guard column (HSS T3, 1.8 µm) at a flow rate of 0.2 mL/min. The mobile

167 

phase and gradient elution program were the same as those used on the LC/ion trap MS system.

168 

Example MRM chromatograms for a standard mixture and two ambient samples are shown in

169 

Fig. 3.

170 

Ambient particulate matter samples. Ambient aerosol samples analyzed in this study

171 

were collected in Hong Kong and in Guangdong province, both are part of the Pearl River

172 

Delta (PRD) in southern China. The sampling dates were synchronized to follow the same 1-

173 

in-6 day schedule in November and December 2010. The sites include one suburban and one

174 

roadside in Hong Kong, and four sites in Guangdong covering suburban, urban, industrial and

175 

residential areas45. PM2.5 samples were collected onto pre-baked 20×25 cm quartz fiber filters

176 

using high-volume (HV) aerosol samplers (Tisch Environmental, Cleves, OH). The sampled

177 

filters were stored in a freezer at –18 ºC until analysis.

178 

An aliquot of 25 cm2 was removed from each HV filter and extracted with high purity

179 

methanol (LiChrosolv LC grade, 99.9%; Merck Millipore) in a sonication bath three times

180 

using 3, 2, and 1 mL in sequence, each extraction for 30 min. Extracts were filtered with a pre-

181 

rinsed polytetrafluoroethylene (PTFE) syringe filter (Bulk Acrodisc CR 13 mm, pore size: 0.25

182 

µm; Pall Life Sciences), combined and reduced in volume to near dryness under high-purity

183 

nitrogen, followed by reconstitution in 100 L 1:1 water: methanol (v/v) containing 200 ppb

184 

D17-octyl sulfate.

185  186 

3. Results and Discussion

187 

Synthesis of organosulfate standards. Two previous studies reported the synthesis of β-

188 

pinene OS35 and limonene OS36 using monoterpene oxide (i.e., β-pinene oxide and limonene

189 

oxide) as the starting materials. This synthetic approach usually resulted in a mixture of OS

190 

regioisomers due to the poor regioselective opening of the epoxide with the sulfate nucleophile

191 

(i.e., structures bp-1 to bp-4 in Fig. 1). In the synthesis of β-pinene OS, Iinuma et al35 did not

192 

attempt to separate the OS mixture products, while in the synthesis of limonene OS36 ion

193 

exchange hydrophilic column used to isolate the limonene OSs from other side products was

194 

not able to separate the isomeric mixture of the limonene OSs. Different from the previous

195 

work, we have developed an efficient and unified route that employs commercially available



 

ACS Paragon Plus Environment

Page 6 of 28

Page 7 of 28

Environmental Science & Technology

196 

and cheap monoterpenes as starting materials. Upjohn dihydroxylation of the monoterpenes as

197 

the first step in the synthesis can typically produce desired diols (2a, 2b in Fig. 2) with excellent

198 

chemo- and diastereoselectivity without other byproducts. For example, for -pinene OS and

199 

freshly synthesized -pinene OS using Upjohn dihydroxylation and sulfation, only one peak

200 

was observed in its HPLC chromatogram (Fig. 4a), and the NMR spectra (Figs. S6 and S7)

201 

confirm the structures of 3a and 3b in Fig. 2. However, diastereomers are produced when the

202 

substrate has poor stereocontrol group (steric hindrance), as in the case of limonene. Two

203 

diastereomers in a 5:2 ratio, which were determined by LC/MS (Fig. 3a & Fig. S10b), were

204 

formed from Upjohn dihydroxylation of limonene and subsequent sulfation,39. To achieve

205 

improved stereoselectivity, we then carried out Sharpless asymmetric dihydroxylation on

206 

limonene using AD mix-β,40 in which the bulky chiral ligand [(DHQD)2-PHAL] bound to OsO4

207 

effectively differentiates the two faces of the alkene (absolute configuration) (Fig. S2), thereby

208 

leading to the production of the diol 2c with excellent diastereoselectivity (1H-NMR cannot

209 

detect the minor isomer, 2c’ in Fig. 2). Upon sulfation, the diastereomeric ratio of limonene

210 

OS (3c: 3c’) was determined to be 20:1 by LC/MS (Fig. S10a).

211 

To achieve regioselective monosulfation and minimize the possible disulfation, one

212 

equivalent of sulfating reagent (SO3·py) 41, 42 was used. Gratifyingly, such regioselectivity was

213 

successfully achieved as SO3·py predominately reacted with the less sterically hindered

214 

hydroxyl group (i.e., the primary alcohol in 2b and the secondary alcohols in 2a and 2c). The

215 

minor regioisomeric OS products were formed, as evidenced by the detection of four limonene

216 

OS peaks when we injected the sulfation products of limonene after Upjohn dihydroxylation

217 

prior to flash chromatography purification. The minor regioisomers were removed in the

218 

subsequent purification step by flash chromatography on silica gel. This two-step method can

219 

be applied to synthesize OSs from other carbon-carbon double bond-containing monoterpenes.

220 

We note that the minor regioisomers, although not isolated in the current synthesis work, could

221 

be formed in the atmosphere and worth future efforts in creating their pure standards in order

222 

to quantify their atmospheric importance.

223 

Two peaks appeared in the HPLC chromatograms of the synthesized standards for the OSs

224 

derived from limonene and limonaketone (Fig. 3a). These two peaks arose from diastereomers.

225 

Diastereomers have been reported to be separated on reverse-phase columns46. For -pinene

226 

OS and freshly synthesized -pinene OS, only one peak was observed in their HPLC

227 

chromatogram, which might be attributed to the excellent diastereoselectivity achieved due to

228 

the presence of the bridgehead carbon.



 

ACS Paragon Plus Environment

Environmental Science & Technology

229 

MS/MS spectra of the synthesized standards: The MS2 spectra for the synthesized OSs

230 

are shown in Figs. S13-S16. Fragments detected in the MS2 spectra are explained in Schemes

231 

S1-S4. The synthesized OSs all show HSO4- at m/z 97 as the dominant fragment ion. Limonene

232 

OS produces a noticeable fragment at m/z 135 (Fig. S13), resulting from loss of H2O and •OSO3

233 

(Scheme S1). -pinene OS shows two relatively abundant fragments at m/z 169 and 151 (Fig.

234 

S14), corresponding to loss of •SO3 and a further loss of H2O, respectively (Scheme S2). -

235 

pinene OS also has a prominent fragment ion •SO3- (m/z 80) (Fig. S15) resulting from the

236 

sulfate function group (Scheme S3). Surratt et al14 reported MS2 spectra of two structural

237 

isomers each for - and -pinene OSs produced in a chamber study (i.e., aP_1/3 and aP_2/4

238 

derived from -pinene and bP_1/3 and bP_2/4 derived from -pinene in Fig. 1). The MS2

239 

spectra of our synthesized OS are mostly consistent with their result and divergences were

240 

minor. For -pinene OS and -pinene OS, we did not attempt to synthesize the other structural

241 

isomers (i.e., aP_2/3/4 and bP_2/3), as they were most likely minor products in our synthesis

242 

strategy and were not targeted for further isolation if some small amounts were produced.

243 

Limonaketone OS produces multiple low-intensity fragment ions, including m/z at 171,

244 

153, 152, 137, 135, and 93 (Fig. S16 and Scheme S4). The fragment characteristics recorded

245 

in this work could be used to assist identification of these OSs in ambient samples, especially

246 

for researchers who do not have access to the standards.

247 

Comparison of LC/MS responses between authentic standards and surrogates. Octyl

248 

sulfate 22-24 and camphor sulfonic acid 13,26-28 are commonly employed as surrogates to quantify

249 

monoterpene-derived OSs because of the similar structures and possessing the same carbon

250 

numbers. 2-ethylhexyl sulfate

251 

surrogates. We established calibration curves for these OSs previously used as quantification

252 

surrogates and for the newly synthesized OSs under both EIC and MRM quantification modes.

253 

Slopes of the calibration curves on the basis of both ppm/ppb and molar concentrations are

254 

summarized in Table 2. The molar slopes provide comparison of MS response per molecular

255 

basis.

25

and dodecyl sulfate

23

are also used in some studies as

The EIC molar slope of the calibration curves ranged from the low value of 0.046

256  257 

0.014 for limonaketone OS to the high value of 0.295

258 

differing by a factor of 6.4 between the two. The two C8 alkyl OSs differ in their MS response

259 

by a factor of 2.3, with the branched C8 OS (i.e., 2-ethylhexyl sulfate) giving a higher response.

260 

These results indicate that the MS response of the [M-H]- ion is not dominated by the sulfate



 

0.070 for 2-ethylhexyl sulfate,

ACS Paragon Plus Environment

Page 8 of 28

Page 9 of 28

Environmental Science & Technology

261 

functional group, instead, the remaining carbon chain structure significantly affects the MS

262 

response, likely through the ESI process.

263 

Under the EIC quantification mode, octyl sulfate, which has a similar carbon number (but

264 

a different carbon skeleton) to the synthetic standards, has a MS response 42% larger than -

265 

pinene OS, 50% larger than -pinene OS, 65% larger than limonene OS, and 171% larger than

266 

limonaketone OS. On the other hand, the molar slopes of camphor sulfonic acid-pinene OS,

267 

-pinene OS, limonene OS are similar, not statistically different at the 95% confidence level

268 

as evaluated using t-test. Camphor sulfonic acid has a similar carbon skeleton as the three

269 

monoterpenes, although it possesses a different polar functional group tail from those of OSs.

270 

The similar MS responses indicate that the carbon skeleton plays a dominating role in affecting

271 

the MS response of [M-H]- ions. As for limonaketone OS, the additional ketone group lowers

272 

its MS response by ~60% in comparison with camphor sulfonic acid and the three monoterpene

273 

OSs, also indicating the carbon chain structure is an influential factor. However, we could not

274 

rule out the possibility that the high water content in mobile phase (water: methanol at 84:26)

275 

at the eluting time of limonaketone OS could have a significant factor contributing to its lower

276 

MS response. The results, although based on a limited number of authentic OSs, suggest that

277 

carbon chain structural similarity is an important selection criterion for surrogates in

278 

quantifying OSs without available standards. Specifically for monoterpene-derived OS,

279 

camphor sulfonic acid is a better quantification surrogate than alkyl OSs in LC/MS analysis

280 

using [M-H]- as the quantification ion. For OSs with additional polar function groups, use of

281 

surrogates not matching the polarity would significantly bias quantification.

282 

LC/MS quantification under MRM mode 24, 28-30 is also used in recent research to quantity

283 

ambient OSs. Quantification by MRM usually provides improved sensitivity, due to lower

284 

baseline that can be attained. In current work, the detection limit under the EIC mode was found

285 

to be higher than the MRM mode by 2 to 10 fold (Tables S2 and S4).

286 

Under MRM mode, a series of MS parameters, among them declustering potential (DP)

287 

and collision energy (CE) being more critical, are varied to achieve maximum fragmentation

288 

ion intensity for the target mass transition. The optimal MS conditions are inherently

289 

compound-specific and dependent on the type of mass transition as well. For the same mass

290 

transition of [M-H]-/97, the optimal CE values for 2-ethylhexyl and octyl OSs were ~10 v

291 

more positive than those for the monoterpene OSs (Table S3). When the MS parameters

292 

optimized for octyl OS were used for MRM quantification, the MS response varied from 0.040

293 

to 0.365 among the OSs, differing by a factor up to ~9. More specifically, the MS responses of



 

ACS Paragon Plus Environment

Environmental Science & Technology

294 

the synthesized OS were only 15-40% of that for octyl OS (Table 2). This means the

295 

concentrations of monoterpene OSs would be underestimated by 2.5-6 times if octyl OS were

296 

used as the quantification surrogate. We also determined the MS responses of the OS

297 

compounds under the MS conditions optimized for -pinene OS. The molar slope of octyl OS

298 

decreased from 0.270 to 0.110 while that of -pinene increased from 0.040 to 0.144 when the

299 

MS parameters were switched from those optimized for octyl OS to those for -pinene OS

300 

(Table 2). The nature of MRM quantification made it difficult to select a suitable surrogate for

301 

quantification of OSs without standard, as demonstrated for the monoterpene OSs.

302 

The MS response for camphor sulfonic acid was 0.377 under its optimized MRM

303 

conditions for mass transition of [M-H]-/80. The monitored fragmentation ion for camphor

304 

sulfonic acid was different from those of OSs. This makes camphor sulfonic acid not a good

305 

surrogate candidate under MRM mode for OSs.

306 

Observation of degradation of -pinene OS. -pinene OS, limonene OS and

307 

limonaketone OS showed neither hydrolysis nor any other side reactions in solid state over two

308 

years, as monitored by NMR analysis. However, -pinene OS degradation was clearly revealed

309 

in our LC/MS analysis of the synthesized standard. Four degradation products containing the

310 

sulfate functional group were observed.

311 

In the LC/MS chromatogram of the newly synthesized -pinene OS standard (Fig. 4),

312 

three peaks were detected, one being the expected -pinene OS and two having a [M-H]- at m/z

313 

231. The m/z 231 peaks are postulated to be formed after loss of a H2O moiety from the -

314 

pinene OS (Compounds 8 and 8’ in Fig. 4). The MS2 spectrum of peak 8 shows a prominent

315 

fragment ion at m/z 163 in addition to ions at m/z 80 and 97 that signify the presence of –SO4

316 

group (Fig. S17). The m/z 163 ion could be rationalized as a result of fragmentation of the [M-

317 

H]- ion through retro-Diels-Alder reaction47. This dehydrated -pinene OS was not present in

318 

the NMR spectrum, therefore it is more likely formed during the LC/MS analysis in the

319 

presence of 0.1% acetic acid in the mobile phase. The hypothesis of acid facilitating the

320 

dehydration finds supporting evidence in the LC/MS chromatogram of a -pinene OS standard

321 

prepared in 1% HCl (Fig. 4b), in which the presence of the dehydrated OSs is significantly

322 

enhanced. 13C-NMR analysis of the -pinene OS standard mixed with 1% HCl confirmed the

323 

presence of C=C bond (characteristic chemical shift at 125 and 133 ppm). Fig. 4b also reveals

324 

a second -pinene OS with [M-H]- at m/z 249 (peak b1’) that was not present in Fig. 4a. We

325 

propose the second -pinene OS could be the regioisomer of the synthesized -pinene OS. The

326 

detailed rational and lines of evidence are presented in Section 12 in SI. 10 

 

ACS Paragon Plus Environment

Page 10 of 28

Page 11 of 28

Environmental Science & Technology

327 

Over the course of our study spanning ~20 months, we injected mixtures of the four

328 

synthesized OSs in different batches of LC/MS analysis. In the LC/MS chromatograms of the

329 

standard mixtures prepared from synthesized compounds ~5-20 months beyond their synthesis,

330 

two -pinene OS (m/z 249) and one additional OS with [M-H]- at m/z 263 were recorded while

331 

the m/z 231 OS disappeared (Fig. 4c). When the standard is over 20 months old, the first -

332 

pinene OS peak (peak 3b) disappeared while the m/z 263 peak remained (Fig. 4d).

333 

The accurate mass data for the m/z 263 by-product by MALDI was determined to be

334 

263.0595 ([M+2Na]+ at m/z 309.0379), corresponding to an elemental formula of C10H15SO6-.

335 

The product ion mass spectrum for the [M-H]- ion at m/z 263 using a collision energy of 30 eV

336 

shows a strong fragment ion at m/z 97 (i.e., [HSO4]- bisulfate ion), the typical fingerprint for

337 

OS. The other structurally informative ion is the m/z 231 fragment ion, corresponding to the

338 

loss of O2 from the [M-H]- ion. The MS data appear to suggest that the two extra O atoms

339 

beyond those in the –OSO3 group could be associated with either –OOH or two –OH groups.

340 

We subsequently carried out NMR analysis of the mixture of -pinene OS and its

341 

degradation product. Both 13C NMR and 1H NMR spectra show the presence of C=C bond in

342 

the degradation product. In the 1H NMR spectrum, the resonance with chemical shift at 5.5

343 

ppm is an indication of alkene double bond (CDCl3; 400 MHz; see Fig. S11). In the 13C NMR,

344 

chemical shifts at 125.1 ppm and 134.3 ppm are typical resonances of SP2-hybrized carbons

345 

(CD3Cl; 400 MHz; see Fig. S12). The 13C NMR spectrum shows two signals resonating in the

346 

range of 60-90 ppm, suggesting two different oxygen-bonded carbons.

347 

Two possible structures in line with the NMR evidence are proposed in Fig. 5 (Structure

348 

5 and 5’). The observation of loss of two O atoms in the MS2 spectrum (Fig. S18) appears to

349 

favor the hydroperoxy structure (i.e., Structure 5). The close RT (24.1 min) in MRM

350 

chromatogram of this degradation product to that of -pinene OS (22.9 and 24.1 min) does not

351 

support the diol structure (Structure 5’) either. The similar polarity of the -pinene OS and its

352 

degradation product was also confirmed by the observation that they could not be separated by

353 

silica gel thin-layer chromatography. If the degradation product were a diol, one would expect

354 

earlier elution than the -pinene OS on the reverse phase column, as is the case with

355 

limonaketone OS (RT = 14.2 min) in comparison with limonene OS (RT = 22.2, 24.5 min). The

356 

extra polarity introduced by one carbonyl group in limonaketone OS shifts its retention time to

357 

much shorter time in comparison with the limonene OS. The presence of another hydroxyl

358 

group would similarly increase polarity and decrease the retention time.48 We used MOPAC

359 

2016 to calculate the dipole moments of the OS compounds. The dipole moments of limonene

11 

 

ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 28

360 

OS and limonaketone OS are 1.466 Debye and 3.205 Debye while those of -pinene OS,

361 

Structures 5 and 5’ are 1.511 Debye, 1.663 Debye, and 3.387 Debye, respectively.

362 

indicated by the dipole moment, the replacement of hydroxyl group with hydroperoxyl group

363 

would maintain polarity and the retention time. In summary, the cumulative evidence supports

364 

the identification of the degradation product as Structure 5. The definitive identification has

365 

yet to be achieved with a purified sample.

49-51

As

366 

We propose in Fig. 6 the possible mechanisms for the formation of OS degradation

367 

products (compounds 3b’, 8/8’, and 5 corresponding to m/z at 249, 231, and 263, respectively)

368 

from the synthesized -pinene OS (compound 3b). The isomerization from compound 3b to

369 

3b’ might take place over time and can be facilitated by acid through cyclic sulfate intermediate

370 

that could be opened by external water nucleophile (path a). Dehydration of the tertiary alcohol

371 

of compound 3b (possibly via cyclic sulfate) leads to compound 8 (path b). The subsequent

372 

allylic oxidation of compound 8 with oxygen via ene-type reaction52 forms compound 5.

373 

Similar reaction process was previously proposed to take place in acetonitrile (CH3CN)

374 

solvent.53,54 The generation of compound 8’ might be arisen from skeletal rearrangement of the

375 

cyclic sulfate intermediate (four-membered ring expansion) followed by β-hydrogen

376 

elimination (path c).

377 

To rationalize the intrigue fact that -pinene OS can undergo various degradation while

378 

other monoterpene OSs do not, we speculate on the basis of comprehensive molecular

379 

structural analysis that the degradation uniquely observed with -pinene OS is a result of

380 

generation of the unstable cyclic sulfate intermediate via intramolecular SN2 substitution of the

381 

tertiary alcohol with the sulfate anion (Fig. 6). On contrast, other monoterpene OSs (i.e., -

382 

pinene OS, limonene OS, and limonaketone OS) are not able to form such cyclic sulfate

383 

because the sulfate anion and the neighboring alcohol (serving as the potential leaving group)

384 

are on the ring that prevent the backside attach of sulfate anion on leaving alcohol (SN2

385 

mechanism). We expect similar degradation could occur to other terminal carbon OSs and

386 

caution is warranted for their quantification in future studies.

387 

For the archived ambient samples collected in 2010, the LC/MS chromatograms show the

388 

presence of -pinene OS peak 3b’ (no peak 3b) and the compound 5, consistent with the

389 

observations with the synthesized -pinene OS. The abundance of compound 5 has a moderate

390 

positive correlation with the -pinene OS (peak 3b & 3b’) (R2 = 0.48, n = 68, Fig. S19). Iinuma

391 

et al. 13 also reported the detection of an OS compound having its [M-H]- at m/z 263 in indoor

392 

aerosol chamber experiments of gas-phase ozonolysis of -pinene in the presence of neutral or

12 

 

ACS Paragon Plus Environment

Page 13 of 28

Environmental Science & Technology

393 

acidic sulfate particles and in ambient aerosol samples collected at a forest site in northeastern

394 

Bavaria, Germany. The RT of this OS was close to the m/z 249 OS on a reverse phase column,

395 

which appears to suggest that the m/z 263 OS likely was the same as the one in our work. They

396 

suggested that the m/z 263 OS was the ring opening product of myrtenol with carbonyl group

397 

and aldehyde group (i.e., Structure 9 in Fig. 5). However, our NMR analysis of the degradation

398 

OS product did not show peaks with chemical shifts in the range of 170 ppm to 200 ppm,

399 

negating the possibility of a carbonyl group as shown in Structure 9.

400 

In the recent ambient samples collected in October and November of 2016, no m/z 263

401 

peak was detected. This result suggests that the peroxy -pinene OS was formed during storage.

402 

Two m/z 249 -pinene OS peaks were present in all the recent ambient samples, and the average

403 

relative abundance of the two is 1.9:1. The two m/z 213 OS peaks were also present in these

404 

recent samples, consistent with the degradation behaviors seen for the synthesized -pinene

405 

OS. However, we could not exclude the possibility of formation of the m/z 213 OS in the

406 

atmosphere.

407 

The observation of -pinene OS degradation indicates attention is needed in

408 

characterizing the stability of OSs during storage. Degradation products induced during storage,

409 

if mistaken as atmospheric degradation product, could mislead interpretation of atmospheric

410 

oxidation mechanism.

411 

Ambient abundance of monoterpene-derived OS in ambient samples. Fig. 3 shows

412 

reconstructed MRM ion chromatograms for m/z 249, 251, and 263 for a standard mixture and

413 

two ambient sample examples, one for an archived sample and one for a recent sample. The

414 

synthesized OS standards were all positively identified in the ambient samples, confirmed by

415 

matching the retention times, [M-H]- ions, and characteristic fragments with those of the

416 

synthetic standards. However, the number of -pinene and limonene OS isomers detected differ

417 

in the recent and the archived samples, as noted in previous section. In the MRM chromatogram

418 

of m/z 249 ion that generate m/z 97 fragment ions (Figs. 3b & 3c), additional peaks were present

419 

besides those associated with -/-pinene and limonene. They are likely OSs products derived

420 

from other monoterpenes or the regioisomers of the synthesized OSs. Surratt et al.14 presented

421 

an m/z 249 EIC chromatogram of an ambient sample collected in southeast US and five peaks

422 

were visually detectable, with four of them identified to be -/-pinene OS.

423 

The four target monoterpene-derived OSs were quantified for ambient PM2.5 samples

424 

collected in November and December in 2010 at six locations in the Pearl River Delta. The

425 

concentration statistics are summarized in Table 3 and in the form of box plots in Fig. S24. The

13 

 

ACS Paragon Plus Environment

Environmental Science & Technology

426 

average concentration of individual OSs was at the level of ~ 0.1-1 ng/m3, with limonene OS

427 

and limonaketone OS being more abundant than -/-pinene OS. The maximum

428 

concentrations were up to a few ng/m3 for individual OSs except for -pinene OS,

429 

concentrations of which was for certain underestimated due to its degradation.

430 

The spatial gradient of OSs and sulfate among the three of the Guangdong sites was

431 

largely indiscernible despite their distinct site characteristics ranging from a park surrounding

432 

to an urban center (Fig. S24). Lack of spatial variation was also observed between the two sites

433 

in Hong Kong, one roadside and one suburban site surrounded by vegetation (Fig. S24).

434 

However, there was a considerable difference between sites in Guangdong and Hong Kong in

435 

that sulfate concentration in Guangdong was approximately twice that in Hong Kong (18.7 vs

436 

9.7 g/m3). In comparison, the spatial gradient in the biogenic OSs was larger than that of

437 

sulfate; with the quantities of the monoterpene OSs in Guangdong sites exceed those in Hong

438 

Kong by ~2.5-5 fold. This result likely suggests nonlinear chemistry in the formation of OSs

439 

and the need for accurate quantification of OSs in order to evaluate how OS formation would

440 

respond to reduction in sulfate.

441 

A few studies in the literature reported the concentration of monoterpene OS (m/z 249)

442 

using quantification surrogates. Kristensen and Glasius26 detected two m/z 249 OS peaks and

443 

reported an average concentration of 0.04 ng/m3 for 17 samples collected in a forest site in

444 

Denmark using camphor sulfonic acid as the quantification surrogate under EIC mode. Ma et

445 

al.24 also reported two m/z 249 OS peaks in samples collected in Shanghai and estimated the

446 

seasonal average concentration from 0.087 ng/m3 in spring to 0.366 ng/m3 in winter using octyl

447 

sulfate as the quantification surrogate under MRM mode. Iinuma et al. 35 synthesized -pinene

448 

OS standard and used it for quantification. They reported a -pinene OS concentration of 23

449 

ng/m3 in one single sample from a forest site of Germany. No other m/z 249 OS peaks were

450 

recorded in the m/z 249 EIC chromatogram shown in their paper. They did not report the purity

451 

of their synthesized standard, nor was degradation reported. It is clear more refined

452 

quantification is needed to evaluate whether the large difference is due to different sources of

453 

the standard or a result of genuine differences in the ambient environments.

454 

We are the first to report identification of limonene and limonaketone OSs in ambient

455 

samples. Limonene OS was more abundant than -/-pinene OS in our samples, with an

456 

average concentration of 1.14 and 0.36 ng/m3 in Guangdong and Hong Kong, respectively.

457 

The concentration of limonaketone OS was on average 0.97 ng/m3 in Guangdong and 0.38

458 

ng/m3 in Hong Kong. The highest detected concentration was 3.83 ng/m3 at the suburban site

14 

 

ACS Paragon Plus Environment

Page 14 of 28

Page 15 of 28

Environmental Science & Technology

459 

on 5 Nov. 2010, coinciding with the highest sulfate concentration (38.0 g/m3) among all the

460 

samples. Considering all the samples (n = 55), the limonaketone OS had a moderate correlation

461 

with sulfate (r = 0.59) while the correlation of limonene OS with sulfate was significantly

462 

weaker (r = 0.29). This possibly reflects that limonaketone OS is a later generation product

463 

derived from limonene oxidation. The abundance ratio of the limonaketone OS to the limonene

464 

OS was on average 0.88 in Guangdong while 1.5 in Hong Kong, consistent with the formation

465 

sequence and Hong Kong is downwind of Guangdong in the months of October and November.

466 

The number of monoterpene OS peaks having [M-H]- at m/z 249 detected in this work

467 

was more than any other studies in the literature that rely on LC/MS for quantification. This

468 

outcome is largely attributable to the availability of authentic standards, which permits

469 

optimization of detection sensitivity under MRM quantification mode and thereby enhanced

470 

detectability. Creating more OS standards of high purity, including those minor regioisomers

471 

not isolated in the current work, is suggested and envisaged invaluable in probing the OS

472 

formation mechanisms in the atmosphere.

473  474 

Supporting Information

475 

Detailed synthesis procedures and LC/MS analytical conditions, NMR spectra, and MS2

476 

spectra of the synthesized organosulfate standards. This material is available free of charge via

477 

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

478  479  480 

Acknowledgements We thank AB Sciex (Hong Kong) for providing us the QTRAP4000 mass spectrometer and

481  482 

funding from the Hong Kong University of Science and Technology (SBI15SC04).

483  484 

References (1) Kanakidou, M.; Seinfeld, J. H.; Pandis, S. N.; Barnes, I.; Dentener, F. J.; Facchini, M. C.;

485  486  487 

Van Dingenen, R.; Ervens, B.; Nenes, A.; Nielsen, C. J.; Swietlicki, E.; Putaud, J. P.; Balkanski, Y.; Fuzzi, S.; Horth, J.; Moortgat, G. K.; Winterhalter, R.; Myhre, C. E. L.; Tsigaridis, K.; Vignati, E.; Stephanou, E. G.; Wilson, J. Organic aerosol and global climate modelling: a

488  489  490  491  492 

review. Atmos. Chem. Phys. 2005, 5, 1053-1123. (2) Hallquist, M.; Wenger, J. C.; Baltensperger, U.; Rudich, Y.; Simpson, D.; Claeys, M.; Dommen, J.; Donahue, N. M.; George, C.; Goldstein, A. H.; Hamilton, J. F.; Herrmann, H.; Hoffmann, T.; Iinuma, Y.; Jang, M.; Jenkin, M. E.; Jimenez, J. L.; Kiendler-Scharr, A.; Maenhaut, W.; McFiggans, G.; Mentel, T. F.; Monod, A.; Prevot, A. S. H.; Seinfeld, J. H.;

493 

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

15 

 

ACS Paragon Plus Environment

Environmental Science & Technology

494  495  496 

organic aerosol: current and emerging issues. Atmos. Chem. Phys. 2009, 9 (14), 5155-5236. (3) Weber, R. J.; Sullivan, A. P.; Peltier, R. E.; Russell, A.; Yan, B.; Zheng, M.; de Gouw, J.; Warneke, C.; Brock, C.; Holloway, J. S.; Atlas, E. L.; Edgerton, E. A study of secondary organic

497 

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

498  499  500 

Atmos. 2007, 112, D13302. (4) Nieminen, T.; Manninen, H. E.; Sihto, S. L.; Yli-Juuti, T.; Mauldin, R. L.; Petaja, T.; Riipinen, I.; Kerminen, V. M.; Kulmala, M. Connection of Sulfuric Acid to Atmospheric

501  502  503 

Nucleation in Boreal Forest. Environ. Sci. Technol. 2009, 43 (13), 4715-4721. (5) Hoyle, C. R.; Boy, M.; Donahue, N. M.; Fry, J. L.; Glasius, M.; Guenther, A.; Hallar, A. G.; Hartz, K. H.; Petters, M. D.; Petaja, T.; Rosenoern, T.; Sullivan, A. P. A review of the

504  505  506 

anthropogenic influence on biogenic secondary organic aerosol. Atmos. Chem. Phys. 2011, 11 (1), 321-343. (6) Czoschke, N. M.; Jang, M.; Kamens, R. M. Effect of acidic seed on biogenic secondary

507  508 

organic aerosol growth. Atmos. Environ. 2003, 37 (30), 4287-4299. (7) Gao, S.; Ng, N. L.; Keywood, M.; Varutbangkul, V.; Bahreini, R.; Nenes, A.; He, J. W.; Yoo,

509 

K. Y.; Beauchamp, J. L.; Hodyss, R. P.; Flagan, R. C.; Seinfeld, J. H. Particle phase acidity and

510  511  512 

oligomer formation in secondary organic aerosol. Environ. Sci. Technol. 2004, 38 (24), 65826589. (8) Gao, S.; Keywood, M.; Ng, N. L.; Surratt, J.; Varutbangkul, V.; Bahreini, R.; Flagan, R. C.;

513 

Seinfeld, J. H. Low-molecular-weight and oligomeric components in secondary organic aerosol

514  515  516 

from the ozonolysis of cycloalkenes and -pinene. J. Phys. Chem. A. 2004, 108 (46), 1014710164. (9) Tolocka, M. P.; Jang, M.; Ginter, J. M.; Cox, F. J.; Kamens, R. M.; Johnston, M. V.

517  518 

Formation of oligomers in secondary organic aerosol. Environ. Sci. Technol. 2004, 38 (5), 1428-1434.

519  520  521  522  523 

(10) Iinuma, Y.; Boge, O.; Gnauk, T.; Herrmann, H. Aerosol-chamber study of the -pinene/O3 reaction: influence of particle acidity on aerosol yields and products. Atmos. Environ. 2004, 38 (5), 761-773. (11) Jang, M.; Czoschke, N. M.; Northcross, A. L.; Cao, G.; Shaof, D. SOA formation from partitioning and heterogeneous reactions: Model study in the presence of inorganic species.

524  525 

Environ. Sci. Technol. 2006, 40 (9), 3013-3022. (12) Kleindienst, T. E.; Edney, E. O.; Lewandowski, M.; Offenberg, J. H.; Jaoui, M. Secondary

526  527  528 

organic carbon and aerosol yields from the irradiations of isoprene and -pinene in the presence of NOx and SO2. Environ. Sci. Technol. 2006, 40 (12), 3807-3812. (13) Iinuma, Y.; Muller, C.; Berndt, T.; Boge, O.; Claeys, M.; Herrmann, H. Evidence for the

529  530  531 

existence of organosulfates from -pinene ozonolysis in ambient secondary organic aerosol. Environ. Sci. Technol. 2007, 41 (19), 6678-6683. (14) Surratt, J. D.; Gomez-Gonzalez, Y.; Chan, A. W. H.; Vermeylen, R.; Shahgholi, M.;

16 

 

ACS Paragon Plus Environment

Page 16 of 28

Page 17 of 28

Environmental Science & Technology

532  533 

Kleindienst, T. E.; Edney, E. O.; Offenberg, J. H.; Lewandowski, M.; Jaoui, M.; Maenhaut, W.; Claeys, M.; Flagan, R. C.; Seinfeld, J. H. Organosulfate formation in biogenic secondary

534  535  536  537 

organic aerosol. J. Phys. Chem. A. 2008, 112 (36), 8345-8378. (15) Hatch, L. E.; Creamean, J. M.; Ault, A. P.; Surratt, J. D.; Chan, M. N.; Seinfeld, J. H.; Edgerton, E. S.; Su, Y. X.; Prather, K. A. Measurements of Isoprene-Derived Organosulfates in Ambient Aerosols by Aerosol Time-of-Flight Mass Spectrometry - Part 1: Single Particle

538  539  540  541 

Atmospheric Observations in Atlanta. Environ. Sci. Technol. 2011, 45 (12), 5105-5111. (16) Hatch, L. E.; Creamean, J. M.; Ault, A. P.; Surratt, J. D.; Chan, M. N.; Seinfeld, J. H.; Edgerton, E. S.; Su, Y. X.; Prather, K. A. Measurements of Isoprene-Derived Organosulfates in Ambient Aerosols by Aerosol Time-of-Flight Mass Spectrometry-Part 2: Temporal Variability

542  543  544  545  546 

and Formation Mechanisms. Environ. Sci. Technol. 2011, 45 (20), 8648-8655. (17) Pye, H. O. T.; Pinder, R. W.; Piletic, I. R.; Xie, Y.; Capps, S. L.; Lin, Y. H.; Surratt, J. D.; Zhang, Z. F.; Gold, A.; Luecken, D. J.; Hutzell, W. T.; Jaoui, M.; Offenberg, J. H.; Kleindienst, T. E.; Lewandowski, M.; Edney, E. O. Epoxide Pathways Improve Model Predictions of Isoprene Markers and Reveal Key Role of Acidity in Aerosol Formation. Environ. Sci. Technol.

547  548  549 

2013, 47 (19), 11056-11064. (18) Lin, Y. H.; Knipping, E. M.; Edgerton, E. S.; Shaw, S. L.; Surratt, J. D. Investigating the influences of SO2 and NH3 levels on isoprene-derived secondary organic aerosol formation

550  551 

using conditional sampling approaches. Atmos. Chem. Phys. 2013, 13 (16), 8457-8470. (19) Friedman, B.; Brophy, P.; Brune, W. H.; Farmer, D. K. Anthropogenic Sulfur Perturbations

552  553  554 

on Biogenic Oxidation: SO2 Additions Impact Gas-Phase OH Oxidation Products of - and Pinene. Environ. Sci. Technol. 2016, 50 (3), 1269-1279. (20) Szmigielski, R. Evidence for C5 organosulfur secondary organic aerosol components from

555  556  557 

in-cloud processing of isoprene: Role of reactive SO4 and SO3 radicals. Atmos. Environ. 2016, 130, 14-22. (21) Carlton, A. G.; Pinder, R. W.; Bhave, P. V.; Pouliot, G. A. To What Extent Can Biogenic

558  559  560  561 

SOA be Controlled? Environ. Sci. Technol. 2010, 44 (9), 3376-3380. (22) Gomez-Gonzalez, Y.; Wang, W.; Vermeylen, R.; Chi, X.; Neirynck, J.; Janssens, I. A.; Maenhaut, W.; Claeys, M. Chemical characterisation of atmospheric aerosols during a 2007 summer field campaign at Brasschaat, Belgium: sources and source processes of biogenic

562  563  564 

secondary organic aerosol. Atmos. Chem. Phys. 2012, 12 (1), 125-138. (23) Kahnt, A.; Behrouzi, S.; Vermeylen, R.; Shalamzari, M. S.; Vercauteren, J.; Roekens, E.; Claeys, M.; Maenhaut, W. One-year study of nitro-organic compounds and their relation to

565  566 

wood burning in PM10 aerosol from a rural site in Belgium. Atmos. Environ. 2013, 81, 561-568. (24) Ma, Y.; Xu, X. K.; Song, W. H.; Geng, F. H.; Wang, L. Seasonal and diurnal variations of

567  568  569 

particulate organosulfates in urban Shanghai, China. Atmos. Environ. 2014, 85, 152-160. (25) Stone, E. A.; Hedman, C. J.; Sheesley, R. J.; Shafer, M. M.; Schauer, J. J. Investigating the chemical nature of humic-like substances (HULIS) in North American atmospheric aerosols

17 

 

ACS Paragon Plus Environment

Environmental Science & Technology

570  571  572 

by liquid chromatography tandem mass spectrometry. Atmos. Environ. 2009, 43 (27), 42054213. (26) Kristensen, K.; Glasius, M. Organosulfates and oxidation products from biogenic

573 

hydrocarbons in fine aerosols from a forest in North West Europe during spring. Atmos Environ.

574  575  576 

2011, 45 (27), 4546-4556. (27) Worton, D. R.; Goldstein, A. H.; Farmer, D. K.; Docherty, K. S.; Jimenez, J. L.; Gilman, J. B.; Kuster, W. C.; de Gouw, J.; Williams, B. J.; Kreisberg, N. M.; Hering, S. V.; Bench, G.;

577  578 

McKay, M.; Kristensen, K.; Glasius, M.; Surratt, J. D.; Seinfeld, J. H. Origins and composition of fine atmospheric carbonaceous aerosol in the Sierra Nevada Mountains, California. Atmos.

579  580 

Chem. Phys. 2011, 11 (19), 10219-10241. (28) He, Q. F.; Ding, X.; Wang, X. M.; Yu, J. Z.; Fu, X. X.; Liu, T. Y.; Zhang, Z.; Xue, J.; Chen,

581  582 

D. H.; Zhong, L. J.; Donahue, N. M. Organosulfates from Pinene and Isoprene over the Pearl River Delta, South China: Seasonal Variation and Implication in Formation Mechanisms.

583  584 

Environ. Sci. Technol. 2014, 48 (16), 9236-9245. (29) Olson, C. N.; Galloway, M. M.; Yu, G.; Hedman, C. J.; Lockett, M. R.; Yoon, T.; Stone,

585 

E. A.; Smith, L. M.; Keutsch, F. N. Hydroxycarboxylic Acid-Derived Organosulfates:

586  587  588 

Synthesis, Stability, and Quantification in Ambient Aerosol. Environ. Sci. Technol. 2011, 45 (15), 6468-6474. (30) Hettiyadura, A. P. S.; Stone, E. A.; Kundu, S.; Baker, Z.; Geddes, E.; Richards, K.;

589 

Humphry, T. Determination of atmospheric organosulfates using HILIC chromatography with

590  591  592 

MS detection. Atmos. Meas. Tech. 2015, 8 (6), 2347-2358. (31) Estillore, A. D.; Hettiyadura, A. P. S.; Qin, Z.; Leckrone, E.; Wombacher, B.; Humphry, T.; Stone, E. A.; Grassian, V. H. Water Uptake and Hygroscopic Growth of Organosulfate

593  594  595 

Aerosol. Environ. Sci. Technol. 2016, 50 (8), 4259-4268. (32) Riva, M.; Budisulistiorini, S. H.; Zhang, Z. F.; Gold, A.; Surratt, J. D. Chemical characterization of secondary organic aerosol constituents from isoprene ozonolysis in the

596  597  598  599  600 

presence of acidic aerosol. Atmos. Environ. 2016, 130, 5-13. (33) Rattanavaraha, W.; Chu, K.; Budisulistiorini, H.; Riva, M.; Lin, Y. H.; Edgerton, E. S.; Baumann, K.; Shaw, S. L.; Guo, H. Y.; King, L.; Weber, R. J.; Neff, M. E.; Stone, E. A.; Offenberg, J. H.; Zhang, Z. F.; Gold, A.; Surratt, J. D. Assessing the impact of anthropogenic pollution on isoprene-derived secondary organic aerosol formation in PM2.5 collected from

601 

the Birmingham, Alabama, ground site during the 2013 Southern Oxidant and Aerosol Study.

602  603  604 

Atmos. Chem. Phys. 2016, 16 (8), 4897-4914. (34) 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,

605  606  607 

S.; Knote, C.; Shaw, S. L.; Zhang, Z.; Gold, A.; Surratt, J. D. Examining the effects of anthropogenic emissions on isoprene-derived secondary organic aerosol formation during the 2013 Southern Oxidant and Aerosol Study (SOAS) at the Look Rock, Tennessee ground site.

18 

 

ACS Paragon Plus Environment

Page 18 of 28

Page 19 of 28

Environmental Science & Technology

608  609  610 

Atmos. Chem. Phys. 2015, 15 (15), 8871-8888. (35) Iinuma, Y.; Boge, O.; Kahnt, A.; Herrmann, H. Laboratory chamber studies on the formation of organosulfates from reactive uptake of monoterpene oxides. Phys. Chem. Chem.

611  612  613  614 

Phys. 2009, 11 (36), 7985-7997. (36) Hansen, A. M. K.; Hong, J.; Raatikainen, T.; Kristensen, K.; Ylisirnio, A.; Virtanen, A.; Petaja, T.; Glasius, M.; Prisle, N. L. Hygroscopic properties and cloud condensation nuclei activation of limonene-derived organosulfates and their mixtures with ammonium sulfate.

615  616  617 

Atmos. Chem. Phys. 2015, 15 (24), 14071-14089. (37) Staudt, S.; Kundu, S.; Lehmler, H. J.; He, X. R.; Cui, T. Q.; Lin, Y. H.; Kristensen, K.; Glasius, M.; Zhang, X. L.; Weber, R. J.; Surratt, J. D.; Stone, E. A. Aromatic organosulfates in

618  619  620 

atmospheric aerosols: Synthesis, characterization, and abundance. Atmos. Environ. 2014, 94, 366-373. (38) Gomes, M. Jr.; Antunes, O. A. C. Upjohn catalytic osmium tetroxide oxidation process:

621  622 

Diastereoselective dihydroxylation of monoterpenes Catalysis Communications 2001, 2, 225– 227.

623  624 

(39) Hao, B.; Gunaratna, M. J., Zhang, M., Weerasekara, S., Seiwald, S. N., Nguyen, V. T., Meier, A., Hua, D. H. Chiral-Substituted Poly-N-vinylpyrrolidinones and Bimetallic

625  626 

Nanoclusters in Catalytic Asymmetric Oxidation Reactions. J. Am. Chem. Soc. 2016, 138, 16839-16848.

627 

(40) Ren, J.; Liu, Y.; Song, L.; Tong, R. Scalable Asymmetric Total Syntheses of (+)-

628  629  630 

Psoracorylifol B and (+)-ent-Psoracorylifol C. Org. Lett. 2014, 16, 2986-2989. (41) Tamaruya, Y.; Suzuki, M.; Kamura, G.; Kanai, M.; Hama, K.; Shimizu, K.; Aoki, J.; Arai, H.; Shibasaki, M. Identifying specific conformations by using a carbohydrate scaffold:

631 

Discovery of subtype-selective LPA-receptor agonists and an antagonist. Angew. Chem. Int.

632  633  634 

Edit. 2004, 43 (21), 2834-2837. (42) Hanessian, S.; Wang, X. T.; Ersmark, K.; DelValle, J. R.; Klegraf, E. Total Synthesis and Structural Revision of the Presumed Aeruginosins 205A and B (vol 11, pg 4232, 2009). Org.

635  636  637 

Lett. 2009, 11 (23), 5566-5566. (43) Rodriguez, R.; Ollivier, C.; Santelli, M. Vitamin D: a concise synthesis of the C19 hydroxylated enyne A-ring, an interesting precursor for the preparation of C19 substituted

638  639 

vitamin D analogues. Tetrahedron Lett. 2004, 45 (11), 2289-2292. (44) Waalboer, D. C. J.; Schaapman, M. C.; van Delft, F. L.; Rutjes, F. P. J. T. High-pressure

640  641  642 

entry into platencin. Angew. Chem. Int. Edit. 2008, 47 (35), 6576-6578. (45) Wang Q.,; Feng, Y.; Huang, X. H. H.; Griffith, S. M.; Zhang, T.; Zhang, Q.; Wu, D., Yu, J. Z. Nonpolar organic compounds as PM2.5 source tracers: Investigation of their sources and

643  644  645 

degradation in the Pearl River Delta, China. J. Geophys. Res-Atmos. 2016, 121, 11862-11879. (46) Cia, D.; Bonhomme B.; Azim, E.-M.; Wada, A.; Doly, M.; Aza ̈ıs-Braesco, V. A reversedphase high-performance liquid chromatographic method to analyze retinal isomers. J.

19 

 

ACS Paragon Plus Environment

Environmental Science & Technology

646  647 

Chromatogr. A 1999, 864, 257–262. (47) Rickborn, B. The Retro-Diels-Alder Reaction. Part I. C-C Dienophiles. Org. React.

648  649  650 

1998, 52, 1-393. (48) Pankow, J. F.; Asher, W. E. SIMPOL.1: a simple group contribution method for predicting vapor pressures and enthalpies of vaporization of multifunctional organic compounds. Atmos.

651  652 

Chem. Phys. 2008, 8 (10), 2773-2796. (49) Szafran, M.; Karelson, M. M.; Katritzky, A. R.; Koput, J.; Zerner, M. C. Reconsideration

653 

of Solvent Effects Calculated by Semiempirical Quantum Chemical Methods. J. Comput.

654  655 

Chem. 1993, 14 (3), 371-377. (50) Demus, D.; Inukai, T. Calculation of molecular, dielectric and optical properties of 4 '-n-

656  657  658 

pentyl-4-cyano-biphenyl (5CB). Liq. Cryst. 1999, 26 (9), 1257-1266. (51) Celle, C.; Suspene, C.; Ternisien, M.; Lenfant, S.; Guerin, D.; Smaali, K.; Lmimouni, K.; Simonato, J. P.; Vuillaume, D. Interface dipole: Effects on threshold voltage and mobility for

659  660 

both amorphous and poly-crystalline organic field effect transistors. Org. Electron. 2014, 15 (3), 729-737.

661  662 

(52) Hoffmann, H. M. R. The Ene Reaction, Angew. Chem. Int. Ed. 1969, 8, 556-577. (53) Capdevielle, P.; Maumy, M. Intervention of Ionic Mechanism in Sensitized Photo-

663  664 

Oxygenation of Alpha-Pinene in Protic Media. Tetrahedron Lett. 1980, 21 (25), 2417-2420. (54) Griesbeck, A. G.; Cho, M. 9-Mesityl-10-methylacridinium: An efficient type II and

665 

electron-transfer photooxygenation catalyst. Org. Lett. 2007, 9 (4), 611-613.

20 

 

ACS Paragon Plus Environment

Page 20 of 28

Page 21 of 28

Environmental Science & Technology

Table 1. Organosulfate compounds synthesized in this study Exact mass by MALDI-MS

249.08

295.0587

295.0561

C10H17O5S-

249.08

295.0587

295.0577

C10H17O5S-

249.08

295.0587

295.0611

5-acetyl-2-hydroxy-2C9H15O6Smethylcyclohexyl sulfate

251.06

297.0379

297.0411

IUPAC name

formula

[M-H]-

-pinene OS

2-hydroxy-2,6,6-trimethylbicyclo[3.1.1]heptan3-yl sulfate

C10H17O5S-

-pinene OS

(2-hydroxy-6,6-dimethylbicyclo[3.1.1]heptan-2yl)methyl sulfate

Limonene OS

2-hydroxy-2-methyl-5(prop-1-en-2yl)cyclohexyl sulfate

Limonaketone OS

21 

 

sodium adduct [M+2Na]+

Common Name

Structure

ACS Paragon Plus Environment

Environmental Science & Technology

Page 22 of 28

Table 2. Slopes of standards and surrogates in this study  Name

2-ethylhexyl sulfate

Octyl sulfate

Camphor sulfonic acid

-pinene OS

-pinene OS

Limonene OS

Limonaketone OS

Dodecyl sulfate

209

209

231

249

249

249

251

265

16.7 0.220 0.036

16.4, 17.4 0.197 0.016

15.8, 17.8 0.199 0.027

0.116

0.034

0.360

26.6 0.119

0.089

0.079

0.080

0.046

0.014

0.136

0.045

Structure

EIC

MRM

Quan ion [M-H](m/z) RT (min) Slope (ppm) Molar Slope (×10-2) RT (min) Slope (ppb), #1a Molar Slope (×10-2)

#1a #2b

0.617

22.6 0.146

0.295

0.070 28.1

0.612

0.030

0.293 0.015 0.365

22.0 0.262 0.038

0.176

13.1 0.035

0.126

0.076

0.015

0.018

27.7 0.229

19.0

0.020

0.870

0.110 0.009 0.270

0.377 /

0.014

23.2

0.285

0.359

0.083

0.123

0.144 0.033 0.040

0.006

22.9, 24.1 0.224

0.065

0.090 0.026 0.095

0.011

8.5

22.2, 24.5 0.252

0.037

0.101 0.015 0.110

14.2

/

0.206c

/

0.246 0.082c 0.063

/ /

0.616

a Slopes obtained under MS parameters optimized using -pinene (except for limonaketone, see note c), expressed as the average and standard deviation of multiple calibrations over a period of 10 months. See Table S3 for specific parameter values. b Slopes obtained under MS parameters optimized using octyl sulfate. See Table S3 for specific parameter values. c Slope obtained under MS parameters optimized using limonaketone OS.

22   

ACS Paragon Plus Environment

Page 23 of 28

Environmental Science & Technology

Table 3. Concentration statistics of organosulfates quantified in this work Species -pinene OS (m/z 249) -pinene OS (m/z 249) Limonene OS (m/z 249) Sum of three monoterpene OSs Limonaketone OS (m/z 251) Ratio of limonene and Limonaketone OS

Concentration in Guangdong sites n=36 a (ng/m3) median max avg 0.41 0.27 2.78 0.19 0.19 0.46 1.14 0.94 3.22 1.74 1.47 6.28

Concentrations in Hong Kong n=19 b (ng/m3) avg median max 0.08 0.08 0.17 0.06 0.05 0.12 0.36 0.25 1.60 0.50 0.45 1.86

0.97

0.54

3.83

0.38

0.37

0.78

0.88

0.48

3.8

1.5

1.1

4.7

a

The samples were collected at four sites in Guangdong in November and December 2010. The average and maximum inorganic sulfate are 18.7 and 38.0 g/m3, respectively. b The samples were collected at two sites in Hong Kong in November and December 2010. The average and maximum inorganic sulfate are 9.7 and 17.6 g/m3, respectively.

23   

ACS Paragon Plus Environment

Environmental Science & Technology

  Figure 1. Chemical structures of organosulfuates derived from -pinene, -pinene, limonene and limonaketone. The structures in solid-line boxes were synthesized in this work. bP_4 was a transformation product of bP_1 observed after a prolonged storage period.    

24   

ACS Paragon Plus Environment

Page 24 of 28

Page 25 of 28

Environmental Science & Technology

Figure 2. Reaction strategy for synthesis of monoterpene derived organosulfates

25   

ACS Paragon Plus Environment

Environmental Science & Technology

Figure 3. Reconstructed MRM ion chromatograms: (a) a standard mixture containing eight organosulfates at 500 ng/mL (b) an archived ambient aerosol sample (NS 20101205), and (c) a recent ambient aerosol sample (UST 20161020 am). The peak labels correspond to the structure labels (if available) in Figs. 2 and 6.

26   

ACS Paragon Plus Environment

Page 26 of 28

Page 27 of 28

Environmental Science & Technology

Figure 4. Reconstructed MRM ion chromatograms for m/z 231, 249, and 263. (a) A newly synthesized -pinene OS standard, (b) A newly synthesized -pinene OS standard prepared in 1% HCl, (c) a standard mixture of -pinene OS, -pinene OS, and limonene OS prepared from synthesized standards more than 5 months old, and (d) a standard mixture of -pinene OS, -pinene OS, and limonene OS prepared from synthesized standards more than 20 months old. The peak labels correspond to the structure labels (if available) in Figs. 2 and 6.

27   

ACS Paragon Plus Environment

Environmental Science & Technology

Figure 5. Three possible structures of the degradation OS products of -pinene OS with [MH]- at m/z 263. Structure 5 is the most likely structure conforming LC/MS and NMR evidence. Structure 5’ conforms with NMR data but not LC/MS data. Structure 9 was proposed by Iinuma et al. (2007)13.

Figure 6. Proposed transformation and degradation pathways and products of the synthesized -pinene OS (compound 3b).

28   

ACS Paragon Plus Environment

Page 28 of 28