Investigation into Photo-induced Auto-oxidation of Polycyclic Aromatic

Dec 27, 2018 - Technol. , Just Accepted Manuscript ... Illuminated samples of naphthalene and anthracene demonstrated growth of several new products w...
0 downloads 0 Views 407KB Size
Subscriber access provided by La Trobe University Library

Environmental Processes

Investigation into Photo-induced Auto-oxidation of Polycyclic Aromatic Hydrocarbons Resulting in Brown Carbon Production John Haynes, Keith E. Miller, and Brian J Majestic Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b05704 • Publication Date (Web): 27 Dec 2018 Downloaded from http://pubs.acs.org on January 2, 2019

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 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 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.

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 25

Environmental Science & Technology

1

Investigation into Photo-induced Auto-oxidation of Polycyclic Aromatic Hydrocarbons

2

Resulting in Brown Carbon Production

3

John P. Haynes1, Keith E. Miller1, Brian J. Majestic1*

4

12190

5

CO 80208, USA

6

*corresponding author, [email protected]

E Iliff Ave, Olin Hall, Department of Chemistry and Biochemistry, University of Denver, Denver,

1 ACS Paragon Plus Environment

Environmental Science & Technology

Page 2 of 25

7

Abstract

8

Brown carbon (BrC) is a collection of oxidized atmospheric aromatic compounds detected worldwide

9

with broad functionality. This multifunctional nature allows BrC to be water soluble, bioavailable, and to

10

demonstrate light absorption at multiple wavelengths. Polycyclic aromatic hydrocarbons (PAH) are major

11

primary products of combustion emissions and have long been known to oxidize in the environment as a

12

component of secondary organic aerosols. In this study we have exposed aqueous PAH suspensions to

13

simulated sunlight to investigate oxidized PAH as BrC precursors. Illuminated samples of naphthalene

14

and anthracene demonstrated growth of several new products with absorptions and oxidation consistent

15

with humic-like substances (HULIS). Reactions of aqueous naphthalene, anthracene, and their oxidized

16

derivatives were found to produce chromatographic and spectroscopic evidence of HULIS formation

17

when exposed to sunlight. The association of oxyradicals with HULIS has implications on human health

18

via lung tissue damage; and its absorption character may add to radiative forcing processes in the

19

atmosphere. Overall product character from naphthalene and anthracene indicate reaction mechanism

20

pathways using oxidized alcohol and quinone as intermediate species.

21

2 ACS Paragon Plus Environment

Page 3 of 25

Environmental Science & Technology

22

INTRODUCTION

23

The atmospheric formation of light absorbing carbonaceous aerosols, collectively known as brown carbon

24

(BrC), is a highly varied and robust subject of many recent studies1-2. The accumulation of these

25

compounds of highly varying composition and absorption3 results in a translucent haze producing a

26

yellow to dark orange color, differentiating it from opaque black carbon species4-6. BrC is an atmospheric

27

phenomenon typically observed in urban and wildfire regions and, as such, is found to evolve from

28

carbon combustion emissions7-8 and materialize within cloud water environments1, 9. The absorbance

29

character of these compounds implicates them with global radiative forcing10-11 and they are associated

30

with respiratory health effects12 due to their significant aromatic composition, similar to polycyclic

31

aromatic hydrocarbons12-14.

32

Polycyclic aromatic hydrocarbons (PAH) are persistent pollutants that can affect natural ecosystems

33

thousands of kilometers from their source14-17. Aerosolized PAH pose a health hazard as their intake is

34

associated with respiratory illness and some are implicated as carcinogens18-22. PAH are produced by the

35

inefficient combustion of hydrocarbon fuel sources, primarily as biomass burning15,

36

exhaust25-27 and industrial complex smokestack byproducts28-29, where PM concentrations of 500 ppm are

37

observed near urban regions23. Due to their extended aromatic structure, PAH have an inherent stability in

38

the atmosphere4,

39

globally15-16, 30.

40

PAH also demonstrate varying levels of photosensitivity as some forms produce oxidation products

41

(oxPAH) upon exposure to light16,

42

oxidations, resulting in hydroxylated groups such as carbonyls, hydroxides, and ethers in varying

43

combinations, all of which have been observed downwind of PAH sources34-37. Naturally occurring OH

44

and other oxidants are implicated with these transformations are gas-phase reactive oxygen species

45

(ROS), including ozone, hydroxyl radicals, and superoxide radicals38-40.

19

23-24,

vehicle

which allows them to persist for hours to weeks and, therefore, be distributed

31-33.

During transport, PAH may go through several of these

3 ACS Paragon Plus Environment

Environmental Science & Technology

Page 4 of 25

46

PAH such as naphthalene, anthracene, and phenanthrene are commonly observed to oxidize into several

47

degradation products indicating the probability of many reaction pathways. Suggested mechanisms in the

48

oxidation and oligomerization of phenolic compounds, for example, describe an array of pathways

49

involving oxygen-centered radical intermediates41-42. Irradiated aromatic alcohols may oxidize through an

50

ozonation43 or excited state carbonyl mechanism44 to obtain characteristics similar to atmospheric humic-

51

like substances (HULIS)45. This may serve as analogous to reactive pathways in similarly oxidized

52

aromatics, like oxPAH. Anthracene observed in atmospheric samples is commonly associated with

53

quinones38,

54

branched oligomers are also observed48.

55

The semi-volatile nature and mass-range of PAH allows for solid-vapor equilibria of many different vapor

56

pressures49. This allows the smaller PAH to exist in vapor form50 and larger PAH to adsorb onto or

57

comprise solid particulate matter (PM)51-52, with the solid PAH cores acting as cloud condensation nuclei

58

(CCN)53. The volatile smaller PAH are also generally more water soluble54 and can therefore partition

59

from the gas phase into the cloud water layer. The combination of PAH vapor-phase partitioning and

60

cloud water leaching from particles will increase dissolved PAH concentrations until equilibria is

61

attained, representing a saturated suspension.

62

Given the structural similarities to phenol, the primary objective of this study is to determine if oxPAH

63

undergo similar photochemistry towards oligomerization within bulk model cloud water systems, which

64

may be an unexplored route to BrC, and to elucidate potential reaction pathways of these oxidations. To

65

achieve this, saturated suspensions of parent chain PAH, specifically naphthalene, anthracene,

66

phenanthrene, and pyrene, are exposed to simulated sunlight, and are analyzed for molecular mass,

67

functional groups, light absorption, polarity, as a proxy for degree of oxidation, and number of reaction

68

products.

46-47

of varying forms, and naphthalene is frequently observed with naphthols33 and larger

69

4 ACS Paragon Plus Environment

Page 5 of 25

Environmental Science & Technology

70

MATERIALS AND METHODS

71

Materials

72

All experiments and sample preparation were performed under a laminar-flow hood using HEPA-filtered

73

air. To reduce cross-contamination in the reactor between reactions, Teflon beaker liners were cleaned

74

through a succession of solvent treatments. This process starts with an acetone rinse followed by an

75

overnight bath of 100% HPLC-grade acetonitrile, then a final overnight bath in 5% trace-metal grade

76

nitric acid including a pre- and post-bath triplicate rinse with 18.2 MΩ purified water. Plastic tubes,

77

bottles, and syringes were prepared using a rinse with 100% acetone and dried before sample addition.

78

Reagents

79

Pure PAH reagents used for these reactions include naphthalene (NAP), anthracene (ANT), phenanthrene

80

(PHE), and pyrene (PYR). Oxidized PAH products used include benzoic acid, phthalic acid, 1-naphthol,

81

1,4-naphthoquinone, 9,10-anthracenediol, 1,4 and 9,10-anthraquinone (Sigma, Fisher). All reagents were

82

used as received without further purification. Ultrapure water was used throughout (ThermoScientific

83

Nanopure, Waltham, MA) and collected at a resistance of 18.2 MΩ.

84

Methods

85

Aqueous suspensions of organic samples were created by adding 100 mg of organic crystals to 200 mL of

86

ultrapure water. These mixtures are then capped, inverted ten times, and stored at room temperature in the

87

dark for 24 hours in order to allow the PAH suspension to achieve equilibrium33, 55-56. The reaction vessels

88

are comprised of 100 mL Teflon liners inside jacketed glass beakers which are temperature controlled by

89

water flow from a chiller pump.

90

During the reaction, samples were stirred at 25°C for 16 hours under a xenon lamp passing through an Air

91

Mass, AM 1.5 Global filter (Oriel Sol 1A, Newport Solar Simulator). Output light consisted of 5.4% UV,

92

54.7% visible, and 39.9% IR resulting in a spectrum equivalent to sunlight, and calibrated for a flux of 5 ACS Paragon Plus Environment

Environmental Science & Technology

Page 6 of 25

93

1550 Wm-2 to simulate one sun at sea level at its zenith. Control samples of identical composition were

94

conducted in the dark by covering an additional reaction vessel with commercially available aluminum

95

foil. Sample aliquots of 4 mL were removed at the beginning and end of the 16 hour reaction period.

96

Sample suspensions following the photoreaction are evaluated for oxidized products and their properties

97

by HPLC, UV-vis, ATR-FTIR, ESI/APCI-MS, and SPE analyses, as described below.

98

Analysis

99

UV-vis (Shimadzu UV-1800) absorption spectra are collected by adding bulk extract to 1 cm quartz

100

cuvettes after baseline calibration using ultrapure water. All samples are analyzed from 220 to 700 nm

101

with maximum absorbance set to 4.0 absorbance units.

102

HPLC (Agilent 1100) analysis was performed using a reversed-phase retention gradient mobile phase

103

method. Separation was obtained using 1 mL/min flow rate, 0.1% TFA:ACN gradient of 90:10 to 0:100

104

over 22.5 min plus hold to 28 min, 100 μL sample injection through a Hydro-RP 250 x 4.6 mm column

105

(Phenomenex), and detection at 254 nm. The chosen column uses a C18 stationary phase on beads with

106

80Å pore size. Therefore each peak represents a compound or set of compounds that match a particular

107

degree of oxidation (i.e, polarity).

108

Functional groups were characterized using attenuated total reflectance Fourier transform infrared

109

spectroscopy (ATR-FTIR, ThermoScientific iS5), analyzed from 700 to 4000 cm-1. The ATR-FTIR

110

analyses of functional groups on light and dark samples are performed on solid residues following

111

aqueous evaporation. The diamond crystal was cleaned and prepared for analysis using a swab of

112

isopropanol and background set after drying. Obtained residues were applied directly to the crystal and

113

firm contact was attained by a torque-controlled plunger.

114

6 ACS Paragon Plus Environment

Page 7 of 25

Environmental Science & Technology

115

RESULTS AND DISCUSSION

116

Photochemistry of Parent PAH

117

Figure 1 presents the UV-vis spectra of the pure PAH parent chains before and after exposure to

118

simulated sunlight. Here, we observe that the illuminated NAP (Fig.1a) UV-vis spectrum demonstrates a

119

dramatic increase in absorption relative to the dark reaction, with the illuminated sample displaying a

120

featureless decay from the UV and throughout the visible range. Illumination of ANT shows almost

121

identical behavior (Fig.1a) compared to NAP, with the dark also maintaining a single absorption peak in

122

the UV region with no absorption in wavelengths higher than 300 nm. Visually, illumination of NAP and

123

ANT result in a transition from clear and colorless to a yellow to orange colored solution.

124 125 126 127 128

Figure 1: UV-vis analysis of saturated suspensions of PAH, (a) naphthalene, in blue, and anthracene, in red, (b) phenanthrene, in green, and pyrene, in purple, in ultrapure water following 16 hour reaction period. Lighted samples are represented in the solid spectra, the dashed spectra represents the samples kept in the dark.

129

The illuminated PHE and PYR (Fig.1b) suspensions create similar, but less intense, discolored solutions,

130

relative to the illuminated NAP and ANT. The increase in absorbance following illumination produces an

131

elevated baseline into the near-UV and visible region of the spectrum. The single peak in the PHE light

132

sample is shown to increase in intensity and shift slightly from 237 to 243 nm, while the PYR sample has

133

similarly shifted its peak at 238 nm in the dark to a shoulder at 253 nm in the light. 7 ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 25

134

To further explore the chemical composition of the illuminated PAH, the HPLC chromatograms in Figure

135

2 display the separations of individual products in the dark and illuminated samples. The NAP (Fig 2a)

136

and ANT (Fig 2b) reactions display significantly more products from irradiation than do PHE and PYR

137

(S1). Illuminated NAP samples demonstrate a wide range of products across the elution gradient starting

138

at retentions of three minutes and displaying an even spread of products until 15 minutes. The most

139

significant peaks have retention times of 8, 11, 13, and 15 minutes which, via their retention times, have

140

been identified as phthalic anhydride, benzoic acid, phthalic acid, and 1-naphthol, respectively.

141 142 143 144 145 146 147

Figure 2: HPLC analysis of post-reaction samples of saturated PAH, (a) NAP and (b) ANT. Lighted samples are represented in the solid spectra, the dashed spectra represents the samples kept in the dark. Numbered peaks are identified products based on retention times of standards dissolved in acetonitrile.1. phthalic anhydride, 2. benzoic acid, 3. phthalic acid, 4. naphthol, 5. naphthalene, 6. phthalic anhydride, 7. benzoic acid, 8. phthalic acid, 9. 1,4-naphthoquinone, 10. 9,10-anthraquinone, 11. anthracene.

148

Illuminated ANT samples also show the creation of robust products across this 3 to 15-minute region and

149

extend up to 23 minutes. Several peaks are shared between the NAP and ANT reactions, including

150

phthalic anhydride, benzoic acid, phthalic acid, and both also contain peaks at 3, 4, and 6 minutes that 8 ACS Paragon Plus Environment

Page 9 of 25

Environmental Science & Technology

151

have not yet been identified. Some key differences in these chromatograms include the introduction of

152

new peaks in the ANT reaction within the 4.5 to 6-minute region, and peaks at 14, 15.5, 18.5, and 20

153

minutes. Among these new peaks, the 14 and 18.5-minute peak products were identified as 1,4-

154

naphthoquinone and 9,10-anthraquinone, respectively. These are compared to the very few peaks

155

obtained from the same reactions kept in the dark.

156

PHE and PYR produced a very small number of peaks in both reactions, which relate to the formation of

157

very few products (Figure S1). Both PHE and PYR share significant peaks at 15 minutes in their

158

illuminated and dark samples, but only illuminated PHE shows a strong peak at 16 minutes. The NAP and

159

ANT in the light yield several products which correlates well with their elevated UV-vis absorptions.

160

Conversely, the PHE and PYR samples demonstrate their lack of responses in UV-vis along with few

161

HPLC products.

162

Figure 2a shows that illuminated NAP samples display robust creation of various new products. These

163

products have a fairly wide mass range of 200 to 800 m/z (LC-MS, Figure S2), suggesting significant

164

growth due likely to a combination of oxidation and oligomerization. The indication of this growth is

165

consistent with similar reaction mechanisms of benzene, ROS, and carbon-centered radicals41,

166

Connecting this information with the UV-vis data in Figure 1 gives the strong suggestion that the curve of

167

diminishing light absorption is likely due to shared light absorption between many different products.

168

This suggestion seems to be further supported considering the PHE and PYR chromatographic data in S1,

169

each displaying very few HPLC peaks in conjunction with a minimal rise in their UV-vis absorption.

170

The combination of chemical characteristics discovered, including featureless UV-vis absorption curve,

171

multitude of products, and increases in molecular mass, is highly indicative of atmospheric oxidized

172

compounds collectively referred to as humic-like substances (HULIS)18. The composition of HULIS is

173

significant in that it contains substantial aromatic and oxygenated functional regions18, and is similar in

57.

9 ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 25

174

nature to terrestrial humic acids, giving them their namesake. This creates an intriguing connection

175

considering that there is a strong correlation between BrC and HULIS observed in field samples4.

176

Method of Formation of the Higher Molecular Mass Compounds

177

To better understand the chemical mechanism towards the more massive oxidized products represented in

178

the LCMS, we applied illumination to some of the discovered products in the ANT and NAP reactions.

179

The UV-vis specta of the illuminated reaction of 1,4 naphthoquinone (Fig. 3a), as well as the 1-naphthol

180

reaction (S3), produces robust absorption increases similar to the illuminated NAP and ANT. Specifically,

181

they all demonstrate a UV-vis decay curve with higher absorptions at shorter wavelengths and vice versa,

182

the only deviation being a broadened shoulder in naphthoquinone. This is in sharp contrast to the

183

complete absence of UV-vis absorption seen from 9,10-anthroquinone samples (Fig. 3b). These results

184

strongly suggest that NAP derivatives, e.g., naphthol and naphthoquinone, are likely intermediate

185

structures in the overall reaction mechanisms. Conversely, the ANT derivatives such as 9,10-

186

anthroquinone are shown to be dormant in sunlight and are therefore considered end products.

187 188 189 190 191

Figure 3: UV-vis spectra of post-reaction samples of oxidized reageants. Solid lines represent samples in the light, dashed lines represent samples in the dark, (a) 1,4naphthoquinone, (b) 9,10-anthraquinone. These spectra are displayed in the 500 ppm concentration in which each reaction was produced. 10 ACS Paragon Plus Environment

Page 11 of 25

Environmental Science & Technology

192

The changes in UV-vis spectra across starting materials follow very closely with the reagents observed to

193

create peaks in the previous HPLC data. UV-vis data (Figure S3) show the lack of spectra response for

194

post-reaction samples for benzoic and phthalic acid. The only change observed between light and dark

195

samples of these simple aromatic acids were a more enhanced absorption near 260 nm in the light sample

196

versus the dark. Interestingly, anthracene-diol (Figure S3) and 9,10-anthraquinone (Fig. 3b) not only

197

show profiles from the light to be essentially unchanged from the dark, but UV-vis spectroscopy shows

198

nearly no absorption at all. Neither the single aromatic acids, e.g., phthalic and benzoic acids, nor the

199

oxidized ANT structure, 9,10-anthraquinone, seem to demonstrate the further oxidation necessary for BrC

200

formation and are therefore considered end points, and do not play a role in HULIS production in the

201

context of photoreactions.

202

By contrast, photo-reactions of naphthoquinone (Fig.3a) and naphthol (Figure S3) produced a

203

significantly greater absorbance profile. These spectra in the dark already contain robust absorption

204

peaks; the illuminated naphthol reaction reproduces the UV-vis absorption decay curves seen with the

205

absorption in the NAP and ANT reactions. The naphthoquinone has an elevated baseline that fits with the

206

reactive PAH curves, however there also includes a pronounced shoulder on this spectra curve.

207

The comparison of the UV-vis and HPLC data for the oxPAH offer a similar perspective to the same

208

comparison of the parent PAH. Product formation for oxPAH reactions is summarized in Figure 4 and

209

Figure S4.

11 ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 25

210 211 212 213

Figure 4: HPLC chromatograms of post-reaction samples of oxidized reageants. Solid lines represent samples in the light, dashed lines represent samples in the dark, (a) 1,4naphthoquinone, (b) 9,10-anthraquinone.

214

The HPLC chromatograms for post-reaction samples for benzoic and phthalic acid (Figure S4) bare a

215

similarity to the 9,10-anthroquinone chromatogram (Figure 4b), suggesting that none of these products

216

are photo-active in sunlight. These simple aromatic acids are shown to produce very few new products in

217

the light. Similarly, anthracene-diol (S4) and 9,10-anthroquinone (Fig.4b) show a product profile from the

218

light to be essentially unchanged from the dark side of the reaction in presenting only one main product.

219

In strong agreement with their UV-vis spectra, photo-reactions of naphthol (Figure S4) and

220

naphthoquinone (Fig.4a) produced a large number of robust products, similar to the parent PAH, NAP

221

and ANT. The major production of naphthol reactions look to start eluting at 11 minutes and continuing a

222

raised baseline until 28 minutes. Naphthoquinone photoreaction shows steady production of several

223

products eluting from about 5 minutes until 20 minutes. In summary, the production of several new peaks

224

in 1,4-naphthoquinone (Fig. 4a) and naphthol (Figure S4) are consistent with their respective UV-vis

225

absorption decay curves, while the lack of photolytic response in 9,10-anthroquinone can be seen in its

226

single significant HPLC peak.

227 12 ACS Paragon Plus Environment

Page 13 of 25

Environmental Science & Technology

228

Aromatic stability and functional groups

229

Reactions of oxPAH demonstrate that polycyclic precursors only further oxidize in structures with one

230

stabilized single-aromatic center, e.g., naphthol and naphthoquinone. While 9,10-anthraquinone, with two

231

stable aromatic centers, shows no oxidation. Among the parent PAH, only the linear species displayed a

232

characteristic photolytic oxidation. Among these oxPAH reagents, only multiple-aromatic structures with

233

oxygen containing groups on their external rings were observed to undergo further oxidation and produce

234

a brown carbon effect. Compounds with oxygen-containing groups on the internal rings did not show a

235

response.

236

The series of reactive PAH presented in this study demonstrate a pattern of predictive oxidizing-active

237

structures and fits well within the set of Clar's rules that predicts the stability of PAH based on Lewis

238

structures58-59. Stated briefly, the overall stability of PAH is based on the number of complete benzene

239

rings, or π-sextet, depicted in an aromatic compound’s structure; where the higher number of sextets in

240

the structure leads to lower energies of the aromatic system, and therefore greater stability and lower

241

reactivity. This is particularly evident in the relatively large brown carbon response from the 3-ring

242

anthracene compound, containing only a single sextet structure, and the lack of response from the 3-ring

243

phenanthrene, containing two sextets.

244

Based on our results and Clar’s rule, we hypothesize that 1,4-anthraquinone, having the oxidized portion

245

on the external rings, may contribute a highly oxidizing potential, given the single stabilized sextet.

246

Figure S5 shows the results of our photochemical reactor on saturated suspensions of 1,4-anthraquione.

247

The UV-vis curve and retention pattern observed in light reactions and several oxidation products

248

displays a pattern consistent with Clar’s rules in producing BrC and HULIS character even though 9,10-

249

anthroquinone, with two sextets, was stable under these conditions.

250

The chemical transformations under sunlight can also be traced by looking at the changes in functional

251

groups. Figure 5 shows the addition of functional groups on NAP residues comparing illuminated to dark 13 ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 25

252

samples. The FTIR spectrum in Figure 5A displays the growth of peaks similar to the other PAH FTIR

253

reactions that display light absorbing characteristics, including naphthoquinone and naphthol (Figure S7),

254

and ANT (Figure S6) whose peaks are less pronounced due to decreased surface area and crystal contact.

255 256 257 258 259 260

Figure 5: ATR-FTIR analysis of NAP products in light and dark conditions. Post-reaction samples are presented for illuminated naphthalene (a) and dark naphthalene (b). Light and dark product analyses are represented by the solid and dashed spectra, respectively. These are performed on a diamond crystal, therefore peak activity in the 2400 cm-1 region is considered background uncertainty.

261

Both NAP (Fig 5b) and ANT (Fig S6) dark samples display single peaks at 3048 cm-1 and a group of

262

several peaks from 1600 to 900 cm-1, correlating to poly-aromatic hydrogen stretches and bends,

263

respectively. One difference with the NAP dark sample, relative to the ANT dark sample, is the small dip

264

centered at 3343 cm-1 which may indicate a hydroxyl hydrogen bend or adsorbed water species.

265

Comparing these spectra with the illuminated samples allows for the determination of sunlight-specific

266

growth. The NAP light spectrum (Fig 5a) shows significant growth of new peaks at 3219, 1717, and large

267

baseline dip from 1360 to 950 cm-1. These would indicate the formation of alcohols, carbonyl groups, and

268

ether bridges, respectively. The ANT light spectrum, in S6, displays new groups with modest absorbance

269

peaks at 1673 and stretch from 1317 to 1175 cm-1, which also correlate to carbonyl and ether groups,

270

respectively. An intriguing difference in functional group character between lighted NAP and ANT

271

samples is the lack of hydroxyl stretches in the 3000-3500 cm-1 region of the ANT spectrum. 14 ACS Paragon Plus Environment

Page 15 of 25

Environmental Science & Technology

272

Figure S7 shows bulk functionality on the remaining photo-active systems. Since only oxidized NAP

273

derivatives are observed to be active samples under light and produced any character related to brown

274

carbon, the FTIR analyses will focus on the structural comparison of those products to known HULIS

275

character, demonstrating a potential step in an overall mechanistic pathway of formation. Naphthol and

276

naphthoquinone (Figure S7a and b) demonstrate the three main features found in oxidized NAP

277

functionality. These peaks of interest are the broad peak at 3300 cm-1, sharp peak at 1700 cm-1, and a

278

pronounced broad peak between 1300-1100 cm-1.

279

Suggested Photo-Reaction Pathway from PAH to HULIS/Brown Carbon

280

The photo-oxidation pathways of NAP and ANT clearly evolved in several directions. Some of these

281

directions lead to simple oxidized end points such as benzoic acid, while others are observed to lead to

282

further, high molecular mass oxidation products, as evidenced on the LCMS. Figure 6 represents the

283

overall collection of information that has come from the NAP and ANT reactions and may act as a

284

precursor for further kinetic and mechanistic studies.

15 ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 25

285 286 287 288 289 290

Figure 6: Preliminary suggested reaction pathways in the production of HULIS from (A) naphthalene and (B) anthracene. Solid arrows represent proposed reactions involving HULIS intermediates and dashed arrows represent formation of end products. FAST and SLOW designations are qualitative and based on the prominence of those compounds in the post-reaction sample. Note that all pathways are elucidated from illuminated samples.

291

While there are likely several reaction pathways in these reactions, the data presented here offer strong

292

evidence for some of the major pathways. Figure 6 displays our proposed preliminary mechanistic

293

pathways toward HULIS from NAP and ANT via photochemistry. The shared presence of produced

294

brown carbon with naphthoquinone character in both NAP and ANT reactions strongly suggests that the

295

conversion of naphthoquinone into brown carbon is likely a faster, favored reaction pathway. Conversely,

296

the lack of brown carbon similar to naphthol reactions in NAP reactions indicates that the conversion of 16 ACS Paragon Plus Environment

Page 17 of 25

Environmental Science & Technology

297

naphthol into brown carbon must be a slower process than naphthoquinone. NAP reactions are indicated

298

as producing naphthol at a faster rate than they produce naphthoquinone. This is based on residual

299

naphthol found in NAP products, which overall demonstrate brown carbon with naphthoquinone

300

character (Figure 2).

301

The ANT proposed pathways are depicted with a nearly opposite trend. The HPLC product data suggest

302

that ANT is likely to produce 1,4-naphthoquinone at a very rapid pace along with its brown carbon

303

products, given our previous estimation that these compounds are converted to brown carbon relatively

304

quickly. The observation that ANT also produces compounds that are retained similar to naphthol and

305

1,4-anthraquinone products further suggests that these are being produced as well, albeit at a slower rate

306

relative to naphthoquinone. Since neither of these structures remains in the ANT sample after the

307

reaction, we must conclude that they are oxidized at the same rate as they are being produced.

308

Since every reaction with a brown carbon response have been observed to produce this only in

309

illuminated samples, the specific impact of sunlight cannot be overstated. The excitation of dissolved

310

oxygen or water molecules in the presence of PAH by UV radiation, producing varying ROS such as OH

311

(as observed by EPR spectroscopy, (Figure S8)), would be a likely component to these reactions. Here,

312

we can suggest some reaction scenarios that may arise with produced hydroxyl (OH), superoxide (Ȯ2¯),

313

and singlet oxygen (1O2) species. Reactions of OH typically involve the homolytic capture of acidic

314

hydrogens to form water and a carbon centered radical. This leaves the hydrocarbon susceptible to attack

315

and will result in a hydroxide group at the site of the attack60. PAH reactions likely use OH to produce

316

significant naphthol, and potentially use elemental oxygen, such as singlet or superoxide to produce

317

quinones via Diels-Alder additions61-62. These are further used as favored intermediates toward the

318

formation of brown carbon. The prominence of 1,4 quinones as active intermediates in these reactions

319

may relate to known phenolic oligomerization mechanisms via semiquinone redox-cycling 63. A potential

320

keto-enol tautomerization to form the semiquinone would be available to 1,4 diketone positions while

321

unavailable to the 9,10 anthraquinone structure, which is shown to be inactive in this study. 17 ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 25

322

HULIS Characteristics

323

The characteristic properties observed in brown carbon production in this study show a prominent

324

similarity to behavior in humic-like substances (HULIS), which are seen as a major component of brown

325

carbon64. HULIS is defined as being structurally similar to humic acids, and as such are presented as a

326

collective of compounds containing significant aromatic and oxidized character18. The propensity for

327

atmospheric HULIS material to evolve from combustion processes65 dictates that the bulk structures will

328

contain aliphatic regions of unburned organic fuels66 and significant aromatic areas formed from

329

inefficient combustion23, 67. HULIS samples observed from field campaigns have demonstrated increased

330

average oxidation states of carbon from -2.0 to -0.468 and contain enhanced oxygen:carbon ratios ranging

331

from 0.69 to 1.069-70 presented as oxygen-containing functional groups. Production of these oxidized

332

groups may be due to oxidation during combustion or down field oxidation by atmospheric oxyradicals71,

333

such as hydroxyl and superoxide radicals.

334

The mass distribution of new products formed from NAP and ANT photo oxidation match the expected

335

mass range of HULIS, specifically in regard to major product groups near 150 m/z and extended mass

336

growth up through 800 and 1000 m/z. Supplementary Figure S2 shows the mass range of saturated

337

aqueous suspensions of PAH and oxPAH reactions. Figures S2a and b represent oxidized naphthalene and

338

anthracene LCMS products; while the peaks are more pronounced in the naphthalene masses than the

339

anthracene, their range and overall spread of masses are very similar. The differences between the

340

naphthol (S2c) and naphthoquinone (S2d) are few but pronounced. They both have a similar baseline

341

spread of products across a range of masses from 200 to 800 m/z, however the naphthol has shown

342

interesting bumps in peaks at the 115 and 301 m/z values.

343

To further confirm the HULIS-like nature of these reactions, analyses of eluted products in a HULIS-

344

specific solid phase extraction (SPE) method5, 18, 72 presents further evidence of this activity. Any products

345

retained using this method are detected by UV-vis and HPLC methods as described above, except that

18 ACS Paragon Plus Environment

Page 19 of 25

Environmental Science & Technology

346

these extractions are evaluated in the methanol matrix used for the elution, and are therefore defined as

347

showing HULIS-like retention.

348

Solid phase extraction analyses of NAP and ANT reactions (Figure S9) demonstrate the presence of

349

products with similar retention characteristics to HULIS, confirming the contribution of HULIS in PAH-

350

generated BrC. The UV-vis analyses for each sample demonstrates a similar absorption curve as the

351

original aqueous lighted sample with a mild shoulder, suggesting a similar composition of HULIS

352

products including some prominence of a major product. These features bear no resemblance to the single

353

peak in their aqueous dark samples; therefore this maintains evidence of various HULIS material with

354

some major products of similar composition. The solid phase elution of HULIS products is further

355

confirmed in the HPLC analyses for NAP and ANT extractions, where significant product peaks are

356

observed and thereby demonstrating the presence of individual HULIS-like products.

357

Environmental Implications

358

The discovery of HULIS character in these reactions introduces new potential hazards for PAH. The

359

formation of groups resembling carboxylic acids, i.e. carbonyl and hydroxy groups, suggests a rationale

360

for an increase in solubility and the potential for chelation and mobilization of previously insoluble

361

nutrients73 and metals found in atmospheric particles74. Complexation reactions of environmental

362

nutrients and metals, such as iron, use water soluble organic species with oxygen-containing groups,

363

carboxylic acids for example, to chelate the typically insoluble ferric iron and act as an aqueous-phase

364

system to bring the iron into solution34, 75. This coincides well with relative soluble iron increases near

365

biomass burning regions76-77, and fits with the expected characteristics of oxPAH including phthalic acid's

366

greater aqueous solubility and carboxylic acids groups. The oxygen-containing groups on HULIS and the

367

products of this study may be able to mobilize cytotoxic metals and oxidizers in a manner similar to

368

siderophore compounds that aid organisms in the uptake of iron78-79. The photolytic connection between

19 ACS Paragon Plus Environment

Environmental Science & Technology

Page 20 of 25

369

PAH and increasing ROS emphasizes an intriguing argument for the continued threat these combustion

370

products can have on the environment during their long range transport.

371

Supporting Information. Figures S1-S9 (PDF)

372

Corresponding Author

373

*Phone (303) 871-2986; email: [email protected].

374

ORCID

375

John P. Haynes: 0000-0002-4454-2093

376

Brian J Majestic: 0000-0003-1860-8738

377

Acknowledgements

378

We thank Drs. Benton Cartledge and Gary Bishop for their help and advice on the installation of the

379

photosimulation reactor components. Dr. Bryan Cowen is thanked for his advice on materials preparation

380

for the reactions involving organic aromatic compounds. We thank Drs. Gareth and Sandra Eaton, as well

381

as Dr. Debbie Mitchell for their help in collecting and interpreting the EPR spectra. This study was

382

supported by NSF award 1342599. The authors declare no competing financial interest.

383

20 ACS Paragon Plus Environment

Page 21 of 25

Environmental Science & Technology

384

References

385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428

1. Laskin, A.; Laskin, J.; Nizkorodov, S. A., Chemistry of Atmospheric Brown Carbon. Chem. Rev. 2015, 115 (10), 4335-4382. 2. Lee, H. J.; Aiona, P. K.; Laskin, A.; Laskin, J.; Nizkorodov, S. A., Effect of Solar Radiation on the Optical Properties and Molecular Composition of Laboratory Proxies of Atmospheric Brown Carbon. Environ. Sci. Technol. 2014, 48 (17), 10217-10226. 3. Washenfelder, R. A.; Attwood, A. R.; Brock, C. A.; Guo, H.; Xu, L.; Weber, R. J.; Ng, N. L.; Allen, H. M.; Ayres, B. R.; Baumann, K.; Cohen, R. C.; Draper, D. C.; Duffey, K. C.; Edgerton, E.; Fry, J. L.; Hu, W. W.; Jimenez, J. L.; Palm, B. B.; Romer, P.; Stone, E. A.; Wooldridge, P. J.; Brown, S. S., Biomass burning dominates brown carbon absorption in the rural southeastern United States. Geophys. Res. Lett. 2015, 42 (2), 653-664. 4. Andreae, M. O.; Gelencser, A., Black carbon or brown carbon? The nature of light-absorbing carbonaceous aerosols. Atmos. Chem. Phys. 2006, 6, 3131-3148. 5. Dinar, E.; Riziq, A. A.; Spindler, C.; Erlick, C.; Kiss, G.; Rudich, Y., The complex refractive index of atmospheric and model humic-like substances (HULIS) retrieved by a cavity ring down aerosol spectrometer (CRD-AS). Faraday Discuss. 2008, 137, 279-295. 6. Decesari, S.; Facchini, M. C.; Matta, E.; Mircea, M.; Fuzzi, S.; Chughtai, A. R.; Smith, D. M., Water soluble organic compounds formed by oxidation of soot. Atmos. Environ. 2002, 36 (11), 1827-1832. 7. Jacobson, M. Z., Effects of biomass burning on climate, accounting for heat and moisture fluxes, black and brown carbon, and cloud absorption effects. J. Geophys. Res.-Atmos. 2014, 119 (14), 89809002. 8. De Haan, D. O.; Hawkins, L. N.; Welsh, H. G.; Pednekar, R.; Casar, J. R.; Pennington, E. A.; de Loera, A.; Jimenez, N. G.; Symons, M. A.; Zauscher, M.; Pajunoja, A.; Caponi, L.; Cazaunau, M.; Formenti, P.; Gratien, A.; Pangui, E.; Doussin, J. F., Brown Carbon Production in Ammonium-or Amine-Containing Aerosol Particles by Reactive Uptake of Methylglyoxal and Photolytic Cloud Cycling. Environ. Sci. Technol. 2017, 51 (13), 7458-7466. 9. Zhao, R.; Lee, A. K. Y.; Huang, L.; Li, X.; Yang, F.; Abbatt, J. P. D., Photochemical processing of aqueous atmospheric brown carbon. Atmos. Chem. Phys. 2015, 15 (11), 6087-6100. 10. Chakrabarty, R. K.; Gyawali, M.; Yatavelli, R. L. N.; Pandey, A.; Watts, A. C.; Knue, J.; Chen, L. W. A.; Pattison, R. R.; Tsibart, A.; Samburova, V.; Moosmuller, H., Brown carbon aerosols from burning of boreal peatlands: microphysical properties, emission factors, and implications for direct radiative forcing. Atmos. Chem. Phys. 2016, 16 (5), 3033-3040. 11. Zhang, Y. Z.; Forrister, H.; Liu, J. M.; Dibb, J.; Anderson, B.; Schwarz, J. P.; Perring, A. E.; Jimenez, J. L.; Campuzano-Jost, P.; Wang, Y. H.; Nenes, A.; Weber, R. J., Top-of-atmosphere radiative forcing affected by brown carbon in the upper troposphere. Nat. Geosci. 2017, 10 (7), 486-+. 12. Lin, Y. H.; Arashiro, M.; Clapp, P. W.; Cui, T. Q.; Sexton, K. G.; Vizuete, W.; Gold, A.; Jaspers, I.; Fry, R. C.; Surratt, J. D., Gene Expression Profiling in Human Lung Cells Exposed to Isoprene-Derived Secondary Organic Aerosol. Environ. Sci. Technol. 2017, 51 (14), 8166-8175. 13. Sun, J. Z.; Zhi, G. R.; Hitzenberger, R.; Chen, Y. J.; Tian, C. G.; Zhang, Y. Y.; Feng, Y. L.; Cheng, M. M.; Zhang, Y. Z.; Cai, J.; Chen, F.; Qiu, Y.; Jiang, Z.; Li, J.; Zhang, G.; Mo, Y., Emission factors and light absorption properties of brown carbon from household coal combustion in China. Atmos. Chem. Phys. 2017, 17 (7), 4769-4780. 14. Lei, Y. L.; Shen, Z. X.; Wang, Q. Y.; Zhang, T.; Cao, J. J.; Sun, J.; Zhang, Q.; Wang, L. Q.; Xu, H. M.; Tian, J.; Wu, J. M., Optical characteristics and source apportionment of brown carbon in winter PM2.5 over Yulin in Northern China. Atmos. Res. 2018, 213, 27-33.

21 ACS Paragon Plus Environment

Environmental Science & Technology

429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474

Page 22 of 25

15. Zhang, Y. X.; Tao, S., Global atmospheric emission inventory of polycyclic aromatic hydrocarbons (PAHs) for 2004. Atmos. Environ. 2009, 43 (4), 812-819. 16. Bamforth, S. M.; Singleton, I., Bioremediation of polycyclic aromatic hydrocarbons: current knowledge and future directions. J. Chem. Technol. Biotechnol. 2005, 80 (7), 723-736. 17. Yunker, M. B.; Snowdon, L. R.; MacDonald, R. W.; Smith, J. N.; Fowler, M. G.; Skibo, D. N.; McLaughlin, F. A.; Danyushevskaya, A. I.; Petrova, V. I.; Ivanov, G. I., Polycyclic aromatic hydrocarbon composition and potential sources for sediment samples from the Beaufort and Barents Seas. Environ. Sci. Technol. 1996, 30 (4), 1310-1320. 18. Graber, E. R.; Rudich, Y., Atmospheric HULIS: How humic-like are they? A comprehensive and critical review. Atmos. Chem. Phys. 2006, 6, 729-753. 19. Kim, K. H.; Jahan, S. A.; Kabir, E.; Brown, R. J. C., A review of airborne polycyclic aromatic hydrocarbons (PAHs) and their human health effects. Environ. Int. 2013, 60, 71-80. 20. Li, J. F.; Dong, H.; Xu, X.; Han, B.; Li, X. G.; Zhu, C. J.; Han, C.; Liu, S. P.; Yang, D. D.; Xu, Q.; Zhang, D. H., Prediction of the bioaccumulation of PAHs in surface sediments of Bohai Sea, China and quantitative assessment of the related toxicity and health risk to humans. Mar. Pollut. Bull. 2016, 104 (12), 92-100. 21. Guerreiro, C. B. B.; Horalek, J.; de Leeuw, F.; Couvidat, F., Benzo(a)pyrene in Europe: Ambient air concentrations, population exposure and health effects. Environ. Pollut. 2016, 214, 657-667. 22. Kumar, B.; Verma, V. K.; Sharma, C. S.; Akolkar, A. B., Estimation of Toxicity Equivalency and Probabilistic Health Risk on Lifetime Daily Intake of Polycyclic Aromatic Hydrocarbons from Urban Residential Soils. Hum. Ecol. Risk Assess. 2015, 21 (2), 434-444. 23. Balachandran, S.; Pachon, J. E.; Lee, S.; Oakes, M. M.; Rastogi, N.; Shi, W. Y.; Tagaris, E.; Yan, B.; Davis, A.; Zhang, X. L.; Weber, R. J.; Mulholland, J. A.; Bergin, M. H.; Zheng, M.; Russell, A. G., Particulate and gas sampling of prescribed fires in South Georgia, USA. Atmos. Environ. 2013, 81, 125-135. 24. Ravindra, K.; Sokhi, R.; Van Grieken, R., Atmospheric polycyclic aromatic hydrocarbons: Source attribution, emission factors and regulation. Atmos. Environ. 2008, 42 (13), 2895-2921. 25. Tsai, P. J.; Shih, T. S.; Chen, H. L.; Lee, W. J.; Lai, C. H.; Liou, S. H., Assessing and predicting the exposures of polycyclic aromatic hydrocarbons (PAHs) and their carcinogenic potencies from vehicle engine exhausts to highway toll station workers. Atmos. Environ. 2004, 38 (2), 333-343. 26. Yang, J.; Ma, S. X.; Gao, B.; Li, X. Y.; Zhang, Y. J.; Cai, J.; Li, M.; Yao, L. A.; Huang, B.; Zheng, M., Single particle mass spectral signatures from vehicle exhaust particles and the source apportionment of on-line PM2.5 by single particle aerosol mass spectrometry. Sci. Total Environ. 2017, 593, 310-318. 27. Liu, P.; Zhang, Y. R.; Wang, L. J.; Tian, B.; Guan, B.; Han, D.; Huang, Z.; Lin, H., Chemical Mechanism of Exhaust Gas Recirculation on Polycyclic Aromatic Hydrocarbons Formation Based on Laser-Induced Fluorescence Measurement. Energy Fuels 2018, 32 (6), 7112-7124. 28. Liberti, L.; Notarnicola, M.; Primerano, R.; Zannetti, P., Air pollution from a large steel factory: Polycyclic aromatic hydrocarbon emissions from coke-oven batteries. J. Air Waste Manage. Assoc. 2006, 56 (3), 255-260. 29. Cetin, B.; Yurdakul, S.; Keles, M.; Celik, I.; Ozturk, F.; Dogan, C., Atmospheric concentrations, distributions and air-soil exchange tendencies of PAHs and PCBs in a heavily industrialized area in Kocaeli, Turkey. Chemosphere 2017, 183, 69-79. 30. Reisen, F.; Arey, J., Atmospheric reactions influence seasonal PAH and nitro-PAH concentrations in the Los Angeles basin. Environ. Sci. Technol. 2005, 39 (1), 64-73. 31. Marques, M.; Mari, M.; Audi-Miro, C.; Sierra, J.; Soler, A.; Nadal, M.; Domingo, J. L., Photodegradation of polycyclic aromatic hydrocarbons in soils under a climate change base scenario. Chemosphere 2016, 148, 495-503.

22 ACS Paragon Plus Environment

Page 23 of 25

475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521

Environmental Science & Technology

32. Shang, J.; Chen, J.; Shen, Z. Y.; Xiao, X. Z.; Yang, H. N.; Wang, Y.; Ruan, A. D., Photochemical degradation of PAHs in estuarine surface water: effects of DOM, salinity, and suspended particulate matter. Environ. Sci. Pollut. Res. 2015, 22 (16), 12389-12398. 33. McConkey, B. J.; Hewitt, L. M.; Dixon, D. G.; Greenberg, B. M., Natural sunlight induced photooxidation of naphthalene in aqueous solution. Water Air Soil Pollut. 2002, 136 (1-4), 347-359. 34. Baba, Y.; Yatagai, T.; Harada, T.; Kawase, Y., Hydroxyl radical generation in the photo-Fenton process: Effects of carboxylic acids on iron redox cycling. Chem. Eng. J. 2015, 277, 229-241. 35. Paris, R.; Desboeufs, K. V., Effect of atmospheric organic complexation on iron-bearing dust solubility. Atmos. Chem. Phys. 2013, 13 (9), 4895-4905. 36. Ringuet, J.; Albinet, A.; Leoz-Garziandia, E.; Budzinski, H.; Villenave, E., Diurnal/nocturnal concentrations and sources of particulate-bound PAHs, OPAHs and NPAHs at traffic and suburban sites in the region of Paris (France). Sci. Total Environ. 2012, 437, 297-305. 37. Perraudin, E.; Budzinski, H.; Villenave, E., Identification and quantification of ozonation products of anthracene and phenanthrene adsorbed on silica particles. Atmos. Environ. 2007, 41 (28), 6005-6017. 38. Arangio, A. M.; Tong, H. J.; Socorro, J.; Poschl, U.; Shiraiwa, M., Quantification of environmentally persistent free radicals and reactive oxygen species in atmospheric aerosol particles. Atmos. Chem. Phys. 2016, 16 (20), 13105-13119. 39. Valavanidis, A.; Fiotakis, K.; Vlachogianni, T., Airborne Particulate Matter and Human Health: Toxicological Assessment and Importance of Size and Composition of Particles for Oxidative Damage and Carcinogenic Mechanisms. J. Environ. Sci. Health Pt. C-Environ. Carcinog. Ecotoxicol. Rev. 2008, 26 (4), 339-362. 40. Fu, P. P.; Xia, Q. S.; Hwang, H. M.; Ray, P. C.; Yu, H. T., Mechanisms of nanotoxicity: Generation of reactive oxygen species. J. Food Drug Anal. 2014, 22 (1), 64-75. 41. Vione, D.; Maurino, V.; Minero, C., Photosensitised humic-like substances (HULIS) formation processes of atmospheric significance: a review. Environ. Sci. Pollut. Res. 2014, 21 (20), 11614-11622. 42. Melo, P. S.; Massarioli, A. P.; Denny, C.; dos Santos, L. F.; Franchin, M.; Pereira, G. E.; Vieira, T.; Rosalen, P. L.; de Alencar, S. M., Winery by-products: Extraction optimization, phenolic composition and cytotoxic evaluation to act as a new source of scavenging of reactive oxygen species. Food Chem. 2015, 181, 160-169. 43. Lesko, T.; Colussi, A. J.; Hoffmann, M. R., Sonochemical decomposition of phenol: Evidence for a synergistic effect of ozone and ultrasound for the elimination of total organic carbon from water. Environ. Sci. Technol. 2006, 40 (21), 6818-6823. 44. Yu, L.; Smith, J.; Laskin, A.; George, K. M.; Anastasio, C.; Laskin, J.; Dillner, A. M.; Zhang, Q., Molecular transformations of phenolic SOA during photochemical aging in the aqueous phase: competition among oligomerization, functionalization, and fragmentation. Atmos. Chem. Phys. 2016, 16 (7), 4511-4527. 45. Ofner, J.; Kruger, H. U.; Grothe, H.; Schmitt-Kopplin, P.; Whitmore, K.; Zetzsch, C., Physicochemical characterization of SOA derived from catechol and guaiacol - a model substance for the aromatic fraction of atmospheric HULIS. Atmos. Chem. Phys. 2011, 11 (1), 1-15. 46. Lin, P.; Yu, J. Z., Generation of Reactive Oxygen Species Mediated by Humic-like Substances in Atmospheric Aerosols. Environ. Sci. Technol. 2011, 45 (24), 10362-10368. 47. McKinney, R. A.; Pruell, R. J.; Burgess, R. M., Ratio of the concentration of anthraquinone to anthracene in coastal marine sediments. Chemosphere 1999, 38 (10), 2415-2430. 48. Chan, A. W. H.; Kautzman, K. E.; Chhabra, P. S.; Surratt, J. D.; Chan, M. N.; Crounse, J. D.; Kurten, A.; Wennberg, P. O.; Flagan, R. C.; Seinfeld, J. H., Secondary organic aerosol formation from photooxidation of naphthalene and alkylnaphthalenes: implications for oxidation of intermediate volatility organic compounds (IVOCs). Atmos. Chem. Phys. 2009, 9 (9), 3049-3060. 23 ACS Paragon Plus Environment

Environmental Science & Technology

522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568

Page 24 of 25

49. Bi, X. H.; Sheng, G. Y.; Peng, P.; Chen, Y. J.; Zhang, Z. Q.; Fu, J. M., Distribution of particulate- and vapor-phase n-alkanes and polycyclic aromatic hydrocarbons in urban atmosphere of Guangzhou, China. Atmos. Environ. 2003, 37 (2), 289-298. 50. Manoli, E.; Kouras, A.; Karagkiozidou, O.; Argyropoulos, G.; Voutsa, D.; Samara, C., Polycyclic aromatic hydrocarbons (PAHs) at traffic and urban background sites of northern Greece: source apportionment of ambient PAH levels and PAH-induced lung cancer risk. Environ. Sci. Pollut. Res. 2016, 23 (4), 3556-3568. 51. Avagyan, R.; Nystrom, R.; Lindgren, R.; Boman, C.; Westerholm, R., Particulate hydroxy-PAH emissions from a residential wood log stove using different fuels and burning conditions. Atmos. Environ. 2016, 140, 1-9. 52. Zhai, Y. B.; Li, P.; Zhu, Y.; Xu, B. B.; Peng, C.; Wang, T. F.; Li, C. T.; Zeng, G. M., Source Apportionment Coupled with Gas/Particle Partitioning Theory and Risk Assessment of Polycyclic Aromatic Hydrocarbons Associated with Size-Segregated Airborne Particulate Matter. Water Air Soil Pollut. 2016, 227 (2). 53. Kasumba, J.; Holmen, B. A., Heterogeneous ozonation reactions of PAHs and fatty acid methyl esters in biodiesel particulate matter. Atmos. Environ. 2018, 175, 15-24. 54. Birks, S. J.; Cho, S.; Taylor, E.; Yi, Y.; Gibson, J. J., Characterizing the PAHs in surface waters and snow in the Athabasca region: Implications for identifying hydrological pathways of atmospheric deposition. Sci. Total Environ. 2017, 603, 570-583. 55. Pronk, G. J.; Heister, K.; Woche, S. K.; Totsche, K. U.; Kogel-Knabner, I., The phenanthrenesorptive interface of an arable topsoil and its particle size fractions. Eur. J. Soil Sci. 2013, 64 (1), 121-130. 56. Laha, S.; Luthy, R. G., EFFECTS OF NONIONIC SURFACTANTS ON THE SOLUBILIZATION AND MINERALIZATION OF PHENANTHRENE IN SOIL-WATER SYSTEMS. Biotechnol. Bioeng. 1992, 40 (11), 13671380. 57. Valavanidis, A.; Vlachogianni, T.; Fiotakis, K., Tobacco Smoke: Involvement of Reactive Oxygen Species and Stable Free Radicals in Mechanisms of Oxidative Damage, Carcinogenesis and Synergistic Effects with Other Respirable Particles. Int. J. Environ. Res. Public Health 2009, 6 (2), 445-462. 58. Sola, M., Forty years of Clar's aromatic pi-sextet rule. Front. Chem. 2013, 1. 59. Clar, E.; Schmidt, W., Correlations between photoelectron and ultraviolet-absorption spectra of polycyclic-hydrocarbons and number of aromatic sextets. Tetrahedron 1975, 31 (18), 2263-2271. 60. Lee, J. Y.; Lane, D. A., Unique products from the reaction of naphthalene with the hydroxyl radical. Atmos. Environ. 2009, 43 (32), 4886-4893. 61. Manceau, M.; Rivaton, A.; Gardette, J. L., Involvement of Singlet Oxygen in the Solid-State Photochemistry of P3HT. Macromol. Rapid Commun. 2008, 29 (22), 1823-1827. 62. Nicolaou, K. C.; Snyder, S. A.; Montagnon, T.; Vassilikogiannakis, G., The Diels-Alder reaction in total synthesis. Angew. Chem.-Int. Edit. 2002, 41 (10), 1668-1698. 63. Cassagnes, L. E.; Perio, P.; Ferry, G.; Moulharat, N.; Antoine, M.; Gayon, R.; Boutin, J. A.; Nepveu, F.; Reybier, K., In cellulo monitoring of quinone reductase activity and reactive oxygen species production during the redox cycling of 1,2 and 1,4 quinones. Free Radic. Biol. Med. 2015, 89, 126-134. 64. Chakrabarty, R. K.; Moosmuller, H.; Chen, L. W. A.; Lewis, K.; Arnott, W. P.; Mazzoleni, C.; Dubey, M. K.; Wold, C. E.; Hao, W. M.; Kreidenweis, S. M., Brown carbon in tar balls from smoldering biomass combustion. Atmos. Chem. Phys. 2010, 10 (13), 6363-6370. 65. Fan, X. J.; Song, J. Z.; Peng, P. A., Temporal variations of the abundance and optical properties of water soluble Humic-Like Substances (HULIS) in PM2.5 at Guangzhou, China. Atmos. Res. 2016, 172, 815. 66. Hoffer, A.; Kiss, G.; Blazso, M.; Gelencser, A., Chemical characterization of humic-like substances (HULIS) formed from a lignin-type precursor in model cloud water. Geophys. Res. Lett. 2004, 31 (6). 24 ACS Paragon Plus Environment

Page 25 of 25

569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609

Environmental Science & Technology

67. Jenkins, B. M.; Jones, A. D.; Turn, S. Q.; Williams, R. B., Emission factors for polycyclic aromatic hydrocarbons from biomass burning. Environ. Sci. Technol. 1996, 30 (8), 2462-2469. 68. Kroll, J. H.; Donahue, N. M.; Jimenez, J. L.; Kessler, S. H.; Canagaratna, M. R.; Wilson, K. R.; Altieri, K. E.; Mazzoleni, L. R.; Wozniak, A. S.; Bluhm, H.; Mysak, E. R.; Smith, J. D.; Kolb, C. E.; Worsnop, D. R., Carbon oxidation state as a metric for describing the chemistry of atmospheric organic aerosol. Nat. Chem. 2011, 3 (2), 133-139. 69. Aiken, A. C.; Decarlo, P. F.; Kroll, J. H.; Worsnop, D. R.; Huffman, J. A.; Docherty, K. S.; Ulbrich, I. M.; Mohr, C.; Kimmel, J. R.; Sueper, D.; Sun, Y.; Zhang, Q.; Trimborn, A.; Northway, M.; Ziemann, P. J.; Canagaratna, M. R.; Onasch, T. B.; Alfarra, M. R.; Prevot, A. S. H.; Dommen, J.; Duplissy, J.; Metzger, A.; Baltensperger, U.; Jimenez, J. L., O/C and OM/OC ratios of primary, secondary, and ambient organic aerosols with high-resolution time-of-flight aerosol mass spectrometry. Environ. Sci. Technol. 2008, 42 (12), 4478-4485. 70. Altieri, K. E.; Seitzinger, S. P.; Carlton, A. G.; Turpin, B. J.; Klein, G. C.; Marshall, A. G., Oligomers formed through in-cloud methylglyoxal reactions: Chemical composition, properties, and mechanisms investigated by ultra-high resolution FT-ICR mass spectrometry. Atmos. Environ. 2008, 42 (7), 14761490. 71. Pillar, E. A.; Camm, R. C.; Guzman, M. I., Catechol Oxidation by Ozone and Hydroxyl Radicals at the Air-Water Interface. Environ. Sci. Technol. 2014, 48 (24), 14352-14360. 72. Varga, B.; Kiss, G.; Ganszky, I.; Gelencser, A.; Krivacsy, Z., Isolation of water-soluble organic matter from atmospheric aerosol. Talanta 2001, 55 (3), 561-572. 73. Nardi, S.; Pizzeghello, D.; Schiavon, M.; Ertani, A., Plant biostimulants: physiological responses induced by protein hydrolyzed-based products and humic substances in plant metabolism. Sci. Agric. 2016, 73 (1), 18-23. 74. Gonzalez, D. H.; Cala, C. K.; Peng, Q. Y.; Paulson, S. E., HULIS Enhancement of Hydroxyl Radical Formation from Fe(II): Kinetics of Fulvic Acid-Fe(II) Complexes in the Presence of Lung Antioxidants. Environ. Sci. Technol. 2017, 51 (13), 7676-7685. 75. Paris, R.; Desboeufs, K. V.; Journet, E., Variability of dust iron solubility in atmospheric waters: Investigation of the role of oxalate organic complexation. Atmos. Environ. 2011, 45 (36), 6510-6517. 76. Mahowald, N. M.; Engelstaedter, S.; Luo, C.; Sealy, A.; Artaxo, P.; Benitez-Nelson, C.; Bonnet, S.; Chen, Y.; Chuang, P. Y.; Cohen, D. D.; Dulac, F.; Herut, B.; Johansen, A. M.; Kubilay, N.; Losno, R.; Maenhaut, W.; Paytan, A.; Prospero, J. A.; Shank, L. M.; Siefert, R. L., Atmospheric Iron Deposition: Global Distribution, Variability, and Human Perturbations. Annu. Rev. Mar. Sci. 2009, 1, 245-278. 77. Paris, R.; Desboeufs, K. V.; Formenti, P.; Nava, S.; Chou, C., Chemical characterisation of iron in dust and biomass burning aerosols during AMMA-SOP0/DABEX: implication for iron solubility. Atmos. Chem. Phys. 2010, 10 (9), 4273-4282. 78. Barbeau, K.; Rue, E. L.; Trick, C. G.; Bruland, K. T.; Butler, A., Photochemical reactivity of siderophores produced by marine heterotrophic bacteria and cyanobacteria based on characteristic Fe(III) binding groups. Limnol. Oceanogr. 2003, 48 (3), 1069-1078. 79. Searle, L. J.; Meric, G.; Porcelli, I.; Sheppard, S. K.; Lucchini, S., Variation in Siderophore Biosynthetic Gene Distribution and Production across Environmental and Faecal Populations of Escherichia coli. PLoS One 2015, 10 (3).

610 611

25 ACS Paragon Plus Environment