Aerosol Brown Carbon from Dark Reactions of ... - ACS Publications

Aerosol Brown Carbon from Dark Reactions of Syringol in Aqueous Aerosol ... (2) Department of Environmental Science and Engineering, Fudan University,...
1 downloads 0 Views 486KB Size
Subscriber access provided by UNIV OF DURHAM

Aerosol Brown Carbon from Dark Reactions of Syringol in Aqueous Aerosol Mimics Jian Xu, Tianqu Cui, Brandon Fowler, Alison Fankhauser, Kai Yang, Jason Douglas Surratt, and V. Faye McNeill ACS Earth Space Chem., Just Accepted Manuscript • DOI: 10.1021/ acsearthspacechem.8b00010 • Publication Date (Web): 18 Apr 2018 Downloaded from http://pubs.acs.org on April 19, 2018

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 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry



Aerosol Brown Carbon from Dark Reactions of Syringol in Aqueous Aerosol



Mimics



Jian Xu,1,2 Tianqu Cui,3 Brandon Fowler,4 Alison Fankhauser,1 Kai Yang,1 Jason D. Surratt,3 and



V. Faye McNeill1,*



(1) Department of Chemical Engineering, Columbia University, New York, NY, USA 10027



(2) Department of Environmental Science and Engineering, Fudan University, 220 Handan

7  8  9  10 

Road, Shanghai 200433, P.R. China (3) Department of Environmental Sciences & Engineering, Gillings School of Global Public Health, University of North Carolina at Chapel Hill, Chapel Hill, NC, USA 27599 (4) Department of Chemistry, Columbia University

11 

*to whom correspondence should be addressed: email: [email protected], phone: +1 (212)

12 

854-2869

13 

Abstract

14 

We performed a laboratory investigation of the chemical processing of syringol, a representative

15 

model phenolic compound emitted from wood burning, in concentrated aqueous salt solutions

16 

mimicking tropospheric aerosol particles. For solutions containing chloride salts, we observed the

17 

formation of light-absorbing organic products (‘brown carbon’), accompanied by a phase

18 

separation, within 10 hours under dark conditions. Products were characterized at the molecular

19 

level using ultra-performance liquid chromatography interfaced to diode array detection and high-

20 

resolution quadrupole time-of-flight mass spectrometry equipped with electrospray ionization

21 

(UPLC/DAD-ESI-HR-Q-TOFMS) and matrix-assisted laser desorption ionization interfaced to

ACS Paragon Plus Environment

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

22 

high-resolution TOFMS (MALDI-TOFMS). The UV-Vis spectra, together with high-resolution

23 

mass spectra results, suggest that syringol can be oxidized by dissolved oxygen, and the presence

24 

of Cl- promotes this reaction. Our results provide new insights into the evolution of aerosol optical

25 

properties during aging, specifically the formation of aerosol brown carbon in biomass burning

26 

plumes.

27 

Keywords

28 

aerosols, particulate matter, brown carbon, atmospheric chemistry, biomass burning

29 

Introduction

30 

Light-absorbing organic aerosol material, or aerosol ‘brown carbon’ (BrC), is an important but

31 

poorly quantified component of atmospheric aerosol emissions from biomass burning (BB).1-4

32 

While at long wavelengths light absorption by BrC is negligible compared to black carbon, its

33 

absorption increases with decreasing wavelength below ~400 nm. Previous ambient observations

34 

of aerosols in areas influenced by BB reported that BrC contributed 20–40% of total aerosol

35 

absorption at 350 nm,5 27% of total aerosol absorption at 404 nm, and almost negligible absorption

36 

at 532 nm.6 The optical properties of BB-derived BrC depend in a complex manner on the

37 

combustion conditions,7 fuel type,3, 8-9 ambient conditions,10 chemical functional groups, 11-12 and

38 

mixing state,6, 13 as well as on chemical aging processes.3, 14-16 For example, light absorption by

39 

primary BrC emissions showed a weaker wavelength dependence than did absorption by BrC

40 

formed via secondary processes.3 Aerosol acidity can also modify the absorptivity of BrC. Lin et

41 

al.17 reported that light-absorbing properties of BrC can be affected by the extraction solvent pH.

42 

Previous studies reported that aromatics, especially polycyclic aromatic hydrocarbon derivatives,

43 

polyphenols, and nitrogen-containing aromatics, are major contributors to BrC light absorption.11,

ACS Paragon Plus Environment

Page 2 of 32

Page 3 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

44 

17

BrC has been associated with extremely low-volatility organic compounds7 and large molecular

45 

weight compounds.18-19 In most cases, fresh BB-derived BrC is expected to lose its light-absorbing

46 

properties within one day due to chemical aging, however, certain BrC fraction are rather resistant

47 

to atmospheric aging.18-20 Additional studies suggested that the dark processing of BB-derived BrC

48 

may change the dominant chromophores from lignin fragments in the fresh emissions11, 21 to

49 

nitroaromatic species in the aged plumes.17

50 

It is becoming clear that BrC plays an important role in climate. Consideration of light absorption

51 

by BrC in models changes calculations of the net direct climate effect of organic carbon aerosols

52 

from scattering to close to zero, especially in areas heavily influenced by BB.22-23 Since BB was

53 

an important source of atmospheric aerosols prior to the industrial revolution, understanding the

54 

optical properties of BB aerosols and how they may evolve after emission is important for

55 

calculations of aerosol radiative forcing.24

56 

Methoxylated phenol compounds, including syringol and guaiacol, are products of lignin

57 

combustion, and have long been used as chemical tracers for BB in atmospheric aerosols.25-26 Field

58 

measurements reported that methoxy-phenols were enriched in the atmospheric aqueous phase,

59 

with concentrations up to three or four times higher than predicted based on Henry’s law.27 Several

60 

recent studies suggested that aqueous phase chemistry of methoxy-phenols resulted in BrC

61 

formation.16,

62 

compounds, including syringol, guaiacol and phenol, serves as an important pathway to form

63 

yellow-brown-colored products with higher molecular weight.31 They found that the IR spectra of

64 

the colored products are similar to that of the humic-like substances (HULIS) found in ambient

65 

organic aerosol. They also reported enhanced formation of colored compounds from precursors

66 

with higher methoxy substitution. Kitanovski and co-workers examined the products from aqueous

28-31

Chang and Thompson showed that the aqueous OH oxidation of phenolic

ACS Paragon Plus Environment

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

67 

photonitration of guaiacol and a yellow-colored solution was observed.30 The reaction products

68 

and intermediates identified, including 4-nitroguaiacol and 4,6-dinitroguaiacol, were also

69 

confirmed in the ambient aerosol samples. In another recent study, vanillin, a guaiacol derivative

70 

and another proxy compound for BB emissions, was reported to form yellowish color and high

71 

molecular weight products at higher concentrations when it undergoes aqueous photolysis.29 The

72 

authors demonstrated that a large portion of the organic mass remained in the particle phase and

73 

non-volatile products were formed. More recently, Yu and coworkers performed experiments of

74 

BrC formation from aqueous reactions of phenol, guaiacol and syringol with two oxidants - the

75 

triplet excited states of an aromatic carbonyl (3C*) and hydroxyl radical. They observed that the

76 

product mixture absorbed light over the visible region and attributed this to conjugated double

77 

bonds formed during polymerization and functionalization. In a follow-up study,32 they

78 

investigated the chemical aging of the BrC products and showed that oligomerization is a major

79 

process in the first 2 hours, while functionalization and fragmentation dominate the continued

80 

aging processes. They proposed that aqueous chemistry of phenolics may contribute to the

81 

formation of extremely low-volatility organic compounds (ELVOCs).

82 

Here we report the formation of light-absorbing organic compounds via reactions of syringol in

83 

the absence of light in concentrated salt solutions containing Cl-. High salt concentrations were

84 

used in order to mimic aqueous atmospheric aerosols.33-34 We observed BrC formation,

85 

accompanied under some conditions by the formation of a precipitate phase also containing light-

86 

absorbing organics. Dimer and oligomer product compounds were identified by matrix-assisted

87 

laser desorption ionization interfaced to high-resolution time-of-flight mass spectrometry

88 

(MALDI-TOFMS), and light-absorbing products were characterized at the molecular level using

89 

ultra-performance liquid chromatography interfaced to diode array detection and high-resolution

ACS Paragon Plus Environment

Page 4 of 32

Page 5 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

90 

quadrupole time-of-flight mass spectrometry (UPLC/DAD-ESI-HR-QTOFMS). The chemistry

91 

reported here potentially represents previously unrecognized pathways for BrC formation from

92 

BB.

93 

Methods

94 

Solution preparation. All commercially available chemicals were used as received without

95 

further purification unless otherwise specified: syringol (Acros Organics; 99 %), guaiacol (Acros

96 

Organics; > 99 %), phenol (Alfa Aesar; > 99 %), NaCl (Fisher Scientific; > 99 %), (NH4)2SO4

97 

(Acros Organics; 99.5 %), NH4NO3 (Acros Organics; > 99 %), NH4Cl (Acros Organics; 99.5 %),

98 

KCl (Sigma Aldrich; > 99 %), NaNO3 (Acros Organics; > 99 %), Na2SO4 (Sigma Aldrich; > 99 %),

99 

CaCl2 (Sigma Aldrich; > 99 %), Fe2(SO4)3 (Sigma Aldrich; 97 %), dichloromethane (Sigma

100 

Aldrich; ≥ 99.8 %). For some experiments, the syringol was purified before use. In an effort to

101 

purify the syringol sample, we prepared a saturated syringol solution (0.2 M). After syringol

102 

crystallized, we collected the crystals and used a desiccator to dry them before use.

103 

Salt solutions were prepared using Millipore water (18.2 MΩꞏcm) in a 100 mL or 250 mL flask at

104 

high, near-saturated concentrations in order to mimic atmospheric aerosol conditions to the extent

105 

possible in a bulk system.33-36 Fresh syringol solution (0.1 M) was made prior to each experiment.

106 

The reaction mixtures were prepared by introducing 0.1 M syringol into salt solution. Mixing time

107 

started count when the reaction mixture was transferred into a stirring (700 rpm), 250 mL Pyrex

108 

glass container. Typical reaction mixtures initially contained 5.1 M NaCl and 1 mM syringol. The

109 

syringol concentration used (1 mM) in this study is higher than the syringol concentrations in

110 

ambient aerosol (0.1-1 µM).25, 27, 37 The upper end of the concentration for total methoxyphenols

111 

can be as high as ~ 30 µM.27 In previous studies, the concentration of syringol employed varied

ACS Paragon Plus Environment

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

112 

from 5 μM to 10 mM.31, 38-39 We used an initial syringol concentration of 1 mM in this study in

113 

order to quantify the formation of light-absorbing products. Results of control experiments

114 

featuring lower initial syringol concentrations (0.1 mM and 0.01 mM) can be found in Figure S2.

115 

Control experiments with other salts and organics were performed as indicated in the text. In order

116 

to investigate the role of dissolved oxygen (O2) in the bulk solution, we bubbled O2 into one

117 

reaction mixture and bubbled N2 into another reaction mixture, and the two processes were done

118 

simultaneously. The air flow rate for N2/O2 was regulated to an adequate amount in order to avoid

119 

any splash of the solution. pH of the aqueous NaCl/syringol reaction mixtures was measured using

120 

an electrochemical pH meter (Fisher) and found to be 5.9-6.3 immediately after mixing and 5.1-

121 

5.5 after 24 h. The pH of the bulk solution was not adjusted throughout the reaction process. The

122 

small decrease in pH that was observed may be due to the formation of uncharacterized organic

123 

acid products. The pH of the 1 mM syringol control solutions was 5.9-6.3 and the pH of the

124 

solutions containing ammonium salts was between 5.0-5.5.

125 

The Pyrex glass container was wrapped with two layers of aluminum foil and all the experiments

126 

were performed under ambient temperature and pressure.

127 

UV-Vis experiments. Absorption spectra of the samples were determined using an HP 8453 UV-

128 

visible Spectrophotometer. The sample was pipetted into an open-top quartz cuvette having an

129 

optical length of 10mm.

130 

MALDI-TOFMS. MALDI-TOFMS was used to detect high molecular weight compounds in

131 

reaction mixtures after stirring 24 hours. The method is similar to that described by Shapiro et al.33

132 

Briefly, 0.4 µL of the mixture and 0.6 µL of the matrix were pre-mixed, spotted on a stainless steel

133 

sample plate, air dried before analyzed using a Bruker ultrafleXtreme MALDI TOF/TOF fitted

ACS Paragon Plus Environment

Page 6 of 32

Page 7 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

134 

with a frequency-tripled Nd:YAG  laser  (355 nm). The matrix solution was either 2,5-

135 

dihydroxybenzoic acid (DHB) or α-cyano-4-hydroxycinnamic acid (CHCA). Signal of the matrix

136 

solution alone was subtracted from the spectrum of the mixture.

137 

UPLC/DAD-ESI-HR-QTOFMS. The reaction mixtures were also analyzed by UPLC/DAD-ESI-

138 

HR-QTOFMS (6520 Series, Agilent) at the University of North Carolina. Phase-separated or

139 

DCM-extracted samples from these reaction mixtures after stirring 24 hours were dried and then

140 

dissolved and diluted by a factor of 15 in a 50:50 (v/v) solvent mixture of high-purity methanol

141 

(LC-MS CHROMASOLV-grade, Sigma-Aldrich) and water (Milli-Q, 18.2 Mꞏcm) prior to

142 

UPLC/DAD-ESI-HR-QTOFMS analysis. In addition, the aqueous phase of the reaction mixtures

143 

was diluted by a factor of 5 in the same solvent mixture used for the phase-separated samples. It

144 

should be noted that these phase-separated and aqueous-phase samples were filtered to remove

145 

any potential insoluble precipitates and transferred to amber UPLC auto-sampler glass vials prior

146 

to UPLC/DAD-ESI-HR-QTOFMS analysis. For comparison to the phase-separated and aqueous-

147 

phase samples obtained from the reaction mixtures, a 1 mM syringol standard prepared in the same

148 

solvent mixture and a solvent blank were also analyzed. Details of the UPLC/DAD-ESI-HR-

149 

QTOFMS operating conditions can be found elsewhere,40-41 and are also summarized in the

150 

Supporting Information. Briefly, the chromatographic separations were performed on a Waters

151 

ACQUITY UPLC HSS (high-strength silica) T3 column (2.1 × 100 mm, 1.8 µm particle size) by

152 

injecting 10 L of each sample. Each sample was analyzed only in the positive ion mode. For the

153 

positive ion mode, samples elute at a flow rate of 0.3 mL min-1 with a solvent mixture of methanol

154 

containing 0.1% ammonium acetate (LC-MS CHROMASOLV-grade, Sigma-Aldrich) and water

155 

containing 0.1% ammonium acetate (LC-MS CHROMASOLV-grade, Sigma-Aldrich), operated

156 

in the high-resolution mode (mass resolving power above 12,000).

ACS Paragon Plus Environment

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

157 

Computational chemistry. Geometry optimizations and vibrational frequency calculations were

158 

performed using Jaguar 9.2 (Schrodinger Inc.) with the Maestro graphical interface in order to

159 

predict and interpret UV-Vis absorption of potential products. Calculations were performed using

160 

density functional theory with the B3LYP or M06-2X functional and the 6-31G or 6-311G** basis

161 

set.34, 42 Calculations took minutes to hours when performed on a workstation with 2 8-core Intel

162 

Xeon E5-2630v3 2.4 GHz processors (Silicon Mechanics).

163 

Results

164 

We observed light-absorbing products form in aqueous solutions containing 5.1 M NaCl and 1

165 

mM syringol within 1 hours of mixing. Because all solutions were covered with two layers of

166 

aluminum foil until analysis, the reactions resulting in light-absorbing molecules are not

167 

photochemical in nature. The light absorption of the products varied with the initial NaCl

168 

concentration in the solution, with higher absorption corresponding to higher initial NaCl

169 

concentration or longer reaction time. The formation of light-absorbing products is accompanied

170 

by a phase separation under some conditions; a precipitate formed when the reaction mixture

171 

contained 1 mM syringol and more than 2 M NaCl.

172 

Light-absorbing products. An aqueous syringol solution containing no salt has a broad absorbance

173 

peak at 270 nm at ambient temperatures. Figure 1a shows that absorption at around 470 nm

174 

develops and increases with time for aqueous mixtures containing 1 mM syringol and 5.1 M NaCl,

175 

indicating the formation of one or more light-absorbing products. Figure 1b shows the time 

176 

dependence of absorption at 470 nm. Error bars shown are the standard deviation in the measured

177 

absorbance. Note that data are shown in Figure 1 for up to 9 hours of reaction time, since precipitate

178 

formed after 10 hours of reaction for these initial conditions, as discussed in the next section.

ACS Paragon Plus Environment

Page 8 of 32

Page 9 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

179 

Figure 2 shows the effect of initial NaCl concentration on the formation of light-absorbing

180 

products. Measurements were made 24 hours after stirring. The absorption at 470 nm increases

181 

with increasing initial NaCl concentration, as shown in Figure 2b. Precipitate formation was

182 

observed after 24 h of reaction for mixtures with initial NaCl concentrations greater than 2 M, as

183 

indicated by the shaded region on Figure 2b.

184 

Several control experiments were performed in order to elucidate the nature of the reaction. Figure

185 

S1a compares the absorption of a control sample containing 1 mM syringol with no salts after 24

186 

h compared to 1 mM syringol in 5.1 M NaCl after 9 h. Absorption increases at 470 nm for the

187 

control solution but to a much lesser extent than the 5.1 M NaCl mixture, indicating an important

188 

role for salts in the BrC formation process.

189 

In order to confirm that the light-absorbing products were formed by syringol and not minor

190 

impurities in the syringol sample, we compare the UV-Vis spectra of the reaction mixtures formed

191 

from both the purified and unpurified samples, as shown in Figure S1c. They exhibited almost the

192 

same absorbance at the UV-Vis wavelengths, demonstrating that impurities play a negligible role

193 

in the formation of the observed light-absorbing products.

194 

Another set of control experiments was performed with reaction mixtures initially consisting of

195 

aqueous solutions of 0.1 mM or 0.01 mM syringol and 5.1 M NaCl. BrC formation was also

196 

observed for these lower, more environmentally relevant initial concentrations of syringol. The

197 

data are shown in Figure S2.

 

ACS Paragon Plus Environment

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 1. UV-VIS absorption of solutions initially containing 1 mM syringol with 5.1 M NaCl as a function of time: (a) time series of UV-VIS spectra and (b) time dependence of absorption at 470 nm. The data points represent at least two measurements and the error bars represent the standard deviation.

  198 

ACS Paragon Plus Environment

Page 10 of 32

Page 11 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

199  200  201  202  203 

  Figure 2. The effect of initial NaCl concentration on formation of light-absorbing products: (a) absorbance as a function of wavelength and (b) absorbance at 470 nm as a function of initial NaCl concentration. All samples contained 1 mM syringol. Measurements were taken after mixing for 24 hours. The shaded area indicate that precipitate was formed.

204  205 

 

206 

 

ACS Paragon Plus Environment

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

207 

In order to test the effect of dissolved O2 on the reaction, we compare the UV-Vis spectra from the

208 

O2-bubbling and N2-bubbling experiments (see Figure S1b). The solution that was bubbled with

209 

N2 exhibited minor light absorption in the visible region after 24 hours. For the solution bubbled

210 

with pure O2, the UV-Vis absorption develops significantly as time proceeds. The color of the

211 

solution changed within one hour, more rapidly than the mixtures that were in equilibrium with

212 

air. These results suggest that dissolved O2 drives the formation of light-absorbing products in the

213 

reaction mixtures.

214 

We also performed control experiments using a number of different salts in order to elucidate the

215 

role of NaCl: NaNO3, Na2SO4, (NH4)2SO4, NH4NO3, NH4Cl, CaCl2 and KCl. The solubility limit

216 

for Na2SO4 is close to 0.7 M, leading us to choose salt concentrations that would yield [Na+] ~ 1.4

217 

M for consistency (i.e., 1.4 M NaNO3). The results are compared in Figure S3 and Figure S4. KCl

218 

promotes the efficient formation of light-absorbing products by syringol, with higher absorbance

219 

at 470 nm than NaCl solutions with similar salt concentrations. A small amount of absorption

220 

(roughly 50% of that for an NaCl solution of similar concentration) was observed at 470 nm for 1

221 

mM syringol in 1.4 M NH4Cl or 1.4 M NaNO3 after 24 h. An increase in absorption at 470 nm

222 

was observed in aqueous solutions of syringol and CaCl2 only after 70 h. Absorption at 470 nm

223 

after 24 h for the (NH4)2SO4, Na2SO4 and NH4NO3 solutions is similar to absorption at 470 nm for

224 

the control solution with no salt. The remainder of our discussion is focused on the NaCl

225 

experiments.

226 

Phase separation. Phase separation was observed for the 1 mM syringol/5.1 M NaCl solution after

227 

10 hours of reaction, leading to an increase in the UV-Vis spectrum baseline due to scattering from

228 

precipitate suspended in the solution. The time of onset for precipitate formation varied with

229 

physical parameters such as the surface-to-volume ratio of the reactor and the rate of mixing, which

ACS Paragon Plus Environment

Page 12 of 32

Page 13 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

230 

likely influenced the kinetics of particle nucleation and growth. In the N2/O2 control experiments

231 

precipitate formed in solutions bubbled with O2 within 5 hours, more rapidly than in the mixtures

232 

that were in equilibrium with air. Precipitate did not form in solutions bubbled with N2.

233 

Figure 3 compares the UV-VIS absorption for the phase-separated samples. Figure S5 shows

234 

images of the control sample initially containing 1 mM syringol in the absence of salts, and the

235 

mixture sample initially containing 1 mM syringol and 5.1 M NaCl. There was no visible color

236 

change in the control sample after stirring for 24 hours (Figure S5a), and no precipitate was

237 

apparent after filtering (Figure S5f). The unfiltered 1 mM syringol/5.1 M NaCl reaction mixture

238 

has amber-color precipitates suspended in the solution after 24 h (Figure S5b). After removal of

239 

the precipitates using a 0.2 μm PTFE filter, the filtered mixture sample still showed a darker color

240 

than the control sample (Figure S5c). The precipitate exhibited a similar color to the unfiltered

241 

mixture solution (Figure S5e). The mass of the precipitate greatly exceeded the initial mass of

242 

organic material used in our experiments, indicating that the precipitate included both salt and

243 

organic material. After washing with Millipore water several times to remove NaCl and water-

244 

soluble organics, the precipitate was then dissolved in dichloromethane (DCM). The DCM

245 

extracted solution was yellowish in color (Figure S5d). The corresponding UV-Vis spectra (Figure

246 

3 & Figure 6) suggest that the light-absorbing products are more soluble in DCM than in water.

247  248 

ACS Paragon Plus Environment

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

249  250  251  252 

Figure 3. The UV-Vis spectra of (a) the 1 mM syringol control, (b) the 1 mM syringol/5.1 M NaCl reaction mixture after 24 h of mixing, (c) the filtrate of the reaction mixture, (d) the DCM extract of the colored precipitate after filtration and washing with water. See text and Figure S3 for details.

253  254 

Surface tension. The surface tension of a solution containing 1 mM syringol and 5.1 M NaCl and

255 

a control solution containing only 1 mM syringol were measured immediately after mixing and

256 

again after 24 hours. No change in surface tension was observed.

257 

Mass spectra characterization of the formed products. MALDI-MS has been used to verify the

258 

presence of large organics with little fragmentation in previous studies.43-44 In this study, two types

259 

of matrix (i.e. DHB and CHCA) were used to test the high molecular weight compounds in the

260 

mixture solution. Though DHB has been frequently used in the detection of high molecular weight

261 

organic compounds,33, 45 it has similar mass-to-charge ratio (m/z) to syringol, which resulted in

262 

interference with the signal from the mixture solution. MALDI-TOFMS mass spectra using CHCA

263 

as matrix are shown in Figure 4.

264 

ACS Paragon Plus Environment

Page 14 of 32

Page 15 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

265  266  267  268 

ACS Earth and Space Chemistry

  Figure 4. MALDI-TOFMS mass spectrum (background subtracted) of the 1 mM syringol/5.1 M NaCl mixture solution after 24 h using CHCA as matrix. Only signals higher than 2 times of the matrix background are shown.

269  270 

Weak signal was observed at m/z 154.08, which can be attributed to syringol (C8H10O3). A cluster

271 

of peaks at m/z 304.1, 305.1, and 306.1 was consistently observed in the spectra. The largest peak

272 

in the cluster appeared at m/z 306.1. This mass-to-charge ratio is consistent with a molecular

273 

formula of C16H18O6. Additional significant peaks appeared at 172.04, 212.03, 234.02, and 385.13

274 

amu. The identities of these species are not confirmed. Although most of the high intensity signal

275 

was distributed in the 200 to 450 amu range, the m/z values of the high molecular weight

276 

compounds in the spectra extends up to approximately 700 amu.

277 

Figure 5 shows the results of UPLC/ESI-HR-QTOFMS analysis. For the solvent blank, the 1 mM

278 

syringol/5.1 M NaCl solution, and the aqueous and DCM extracts, significant signal appeared in

279 

the base peak chromatogram (BPC) after ~ 10.3 minutes elution time. The mass spectrum is shown

280 

as a function of m/z in the inset. Syringol is observed at m/z 155 and 177 as the M + H+ and M +

ACS Paragon Plus Environment

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

281 

Na+ ions, respectively. Multiple peaks are observed between 300 and 1000 amu, which is

282 

consistent with the MALDI-TOFMS results. Signal is observed at m/z 305.1 in the product mixture,

283 

consistent with the protonated form of a dimer product with mass 304 amu (C16H16O6). Extracted

284 

ion chromatrogram (EIC) analysis shows that this peak is present in both the aqueous and DCM

285 

extracts of the product mixture, with more abundance in the aqueous extract.

286 

Figure 6 shows the results of UPLC/DAD analysis at 470 nm. These UPLC/DAD chromatograms

287 

have been corrected for dilution and background subtraction. For the solvent blank and the 1 mM

288 

syringol control (in the absence of NaCl) solutions, no significant chromatographic peaks were

289 

observed except for a small peak eluting at 10.12 minutes for the syringol control. Notably, three

290 

substantial light-absorbing compounds at 470 nm were found to elute off the reverse-phase UPLC

291 

column at 7.04, 9.42 and 10.06 minutes for the DCM extracts. Since the compound eluting at 10.06

292 

min has a larger chromatographic peak area than the compound eluting at 10.12 minutes for the

293 

syringol control sample, this slight retention time shift is likely explained by the difference in

294 

abundance, and thus, are the same compound. As shown in Figure S6, the two earlier eluting light-

295 

absorbing species at 7.04 and 9.42 minutes observed from the DCM extracts could not be observed

296 

with the same retention times in the corresponding UPLC/ESI-HR-QTOFMS base peak ion

297 

chromatogram (BPC). A single product, which had a mass of 191.1588 amu (C9H21NO3), in the

298 

DCM extract could be observed at 7.57 minutes in the UPLC/ESI-HR-QTOFMS BPC for the

299 

DCM extracts but not in the UPLC/DAD chromatograms at 470 nm. Since this DCM-extracted

300 

product with mass 191.1588 amu does not overlap with the three compounds observed in Figure

301 

6, the mass 191.1588 amu product does not contribute to the light-absorption properties of the

302 

sample. For the aqueous extracts, only one distinct peak was observed to elute in the UPLC/DAD

303 

chromatogram at 10.12 minutes, which corresponds exactly to the dimer product with mass 304

ACS Paragon Plus Environment

Page 16 of 32

Page 17 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

304 

amu (C16H16O6) detected by the UPLC/ESI-HR-QTOFMS technique. We note that a small peak

305 

was observed at the same retention time for the syringol control solutions, suggesting that small

306 

amounts of the C16H16O6 are present in those solutions. Since the UPLC/DAD and UPLC/ESI-HR-

307 

QTOFMS share the same retention time for this compound, this strongly supports that the dimer

308 

product with mass 304 amu (C16H16O6) is contributing to the light-absorption of the aqueous and

309 

DCM extracts of these samples. However, in comparison to the aqueous extracts, the DCM extracts

310 

yields two additional light-absorption products resolved by the UPLC/DAD at retention times 7.04

311 

and 9.42 minutes that are not observed by UPLC/ESI-HR-QTOFMS. This suggests that alternative

312 

ionization techniques to ESI employed here may be required in the future in order to detect and

313 

characterize them by mass spectrometry. Since these two light-absorbing products elute earlier on

314 

the reverse-phase UPLC column than the light-absorbing dimer product with mass 304 amu, this

315 

suggests they are more hydrophilic (or smaller, possibly monomeric) in nature.

316 

ACS Paragon Plus Environment

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

317 

 

318  319  320  321  322  323  324  325  326  327 

Page 18 of 32

   Figure 5. UPLC/ESI-HR-QTOFMS extracted ion chromatograms (EICs) of m/z 305 for different solutions after stirring 24 hours. ‘Solvent blank’ refers to 50:50 (v/v) solvent mixture of highpurity methanol and water (Milli-Q, 18.2 Mꞏcm) used for preparing samples for UPLC/ ESI-HRQTOFMS analysis. ‘Syringol control’ indicates an aqueous solution of 1 mM syringol in the absence of salt. ‘Water extracts’ refers to the water soluble fraction of the 1 mM syringol/5.1 M NaCl reaction mixture. ‘DCM extracts’ refers to the DCM extracted fraction of the 1 mM syringol/5.1 M NaCl reaction mixture. The inset figure shows the average mass spectra for the major chromatographic peaks eluting between 10.25 and 10.38 min, corresponding to the light purple shading area.

328 

ACS Paragon Plus Environment

Page 19 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

329 

ACS Earth and Space Chemistry

 

330  331  332  333  334  335  336  337  338  339  340  341  342 

Figure 6. UPLC/DAD chromatograms at 470 nm, which are corrected by background subtraction and dilution, for different solutions after stirring 24 hours. Note that these full UV-Vis spectra and UPLC/ESI-HR-QTOFMS mass spectra for these chromatophilic peaks shown here are provided in Figures S8-S12. ‘Solvent blank’ refers to 50:50 (v/v) solvent mixture of high-purity methanol and water (Milli-Q, 18.2 Mꞏcm) used for preparing samples for UPLC/DAD-ESI-HR-QTOFMS analysis. ‘Syringol control’ indicates an aqueous solution of 1 mM syringol in the absence of salt. ‘Water extracts’ refers to the water soluble fraction of the 1 mM syringol/5.1 M NaCl reaction mixture. ‘DCM extracts’ refers to the DCM extracted fraction of the 1 mM syringol/5.1 M NaCl reaction mixture. Light-absorbing products at 470 nm are observed to elute from the reverse-phase UPLC column for the DCM extracts at 7.04, 9.42, and 10.06 minutes. Light-absorbing products are also observed to elute at 10.12 and 10.15 minutes for the syringol control and the aqueous extracts, respectively.

ACS Paragon Plus Environment

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

343 

Discussion

344 

The UV-Vis absorption spectra and the UPLC/DAD chromatograms indicate multiple light-

345 

absorbing organic products (brown carbon) formed in the syringol/NaCl reaction mixtures.

346 

Although we know that the reactions taking place are not photochemical since all samples were

347 

shielded from light, the measured absorbance at 470 nm in our study is similar to the visible

348 

absorption spectra of products observed from photochemical reactions of syringol.16, 31 Due to the

349 

critical role of dissolved O2 that we infer from our control experiments, an oxidative process is

350 

likely.

351 

Methoxy-phenols (guaiacol and catechol) are known to undergo oxidation under dark conditions

352 

in aqueous solutions in the presence of transition metal ions (TMI), forming unstable quinone

353 

species which absorb at around 470 nm.46-49 We investigated the possibility that trace amounts of

354 

TMI contamination in our salts could be causing the observed reactions and found that it is not

355 

likely. All of the high-purity salts used in this study, including ammonium sulfate and other salts

356 

observed to result in no reaction, are stated to contain at most a few ppm of heavy metals including

357 

Fe, corresponding to roughly 10-5 M (see Supporting Information for full details). Further, a

358 

mixture of 5×10-6 M Fe2(SO4)3, 1 mM syringol and 3.1 M (NH4)2SO4 yielded no BrC after 24

359 

hours. Finally, control experiments involving solutions prepared with 1 mM phenol or 1mM

360 

guaiacol and 5.1 M NaCl (Figure S7) did not yield light-absorbing products (recall that TMI-

361 

catalyzed guaiacol oxidation is known to yield BrC46-49). Based on this evidence, we conclude that

362 

the observed reaction is not due to catalysis by trace TMI contaminants.

363 

Next, we discuss the possible role of the salt in the oxidation of syringol. Light-absorbing products

364 

were formed in varying amounts from aqueous solutions of syringol with NaCl, KCl, CaCl2, and

ACS Paragon Plus Environment

Page 20 of 32

Page 21 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

365 

NH4Cl. No reaction was observed with (NH4)2SO4, Na2SO4 and NH4NO3 solutions. The

366 

implication is that chloride promotes the formation of light-absorbing products (although we note

367 

that small amounts of BrC were also formed in aqueous syringol/NaNO3 solutions). Chloride salts

368 

are known to decrease the solubility of O2 in aqueous solutions,50 which should inhibit oxidation,

369 

rather than promote it. Therefore, we consider the interaction of the chloride salts with the organics

370 

in solution.

371 

Phenol is known to ‘salt out’ (in other words, its solubility decreases with increasing salt

372 

concentration, with different salts reducing solubility to a different degree, as quantified by the

373 

Setschenow constant, Ks).51-52 To our knowledge, this phenomenon has not been investigated for

374 

syringol, but we expect its salting-out behavior to be similar to phenol due to their structural

375 

similarities. While our experiments were not performed near the solubility limit for syringol, on a

376 

molecular level, reduced solubility, corresponding to more hydrophobic behavior,53 may lead to

377 

an increased encounter rate among the organics by increasing their local concentration. The trend

378 

among the chloride salts (i.e., BrC production from KCl > NaCl > NH4Cl > CaCl2) is consistent

379 

with an inverse relationship to the expected Setschenow constants of these salts for syringol. For

380 

phenol, Ks,CaCl2 > Ks,NaCl > Ks, KCl.51 Ks,NH4Cl is often similar to Ks,NaCl.54 In other words, maximum

381 

BrC formation is observed when Cl- is present, but salting-out (i.e., the hydrophobic effect) is

382 

minimized.

383 

We tentatively propose that, in the presence of chloride salts and dissolved O2, syringol may be

384 

oxidized by O2 to form a quinone intermediate, which then reacts to form a dimer with a molecular

385 

formula of C16H16O6 and molecular weight of 304.1 amu. Ionic liquid solvents have been observed

386 

to enhance syringol oxidation by TMI catalysts, by stabilizing ionic reactive intermediates via

387 

clustering interactions.55-56 While the fundamental chemistry of this system is quite different, it is

ACS Paragon Plus Environment

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

388 

possible that the role of Cl- in this reactive system is to stabilize reactive intermediates. The

389 

clustering behavior of Cl- is expected to differ from that of sulfate or nitrate due to steric

390 

considerations. The fact that Cl- promoted reactions were not observed for phenol and guaiacol

391 

(Figure S7), may be due to the fact that those species have fewer electron donating functional

392 

groups, and therefore they and intermediates derived from them are expected to be both less

393 

reactive than syringol and less capable of clustering with Cl-.

394 

Results of ab initio simulations indicate that the quinone dimer is likely to absorb at visible

395 

wavelengths. The energies of the highest occupied molecular orbital (HOMO) and lowest

396 

unoccupied molecular orbital (LUMO) were calculated using Jaguar, then the absorption

397 

wavelength was derived from the HOMO-LUMO energy difference using Planck’s equation. The

398 

energy calculations were performed using either the B3LYP functional or M06-2X, and either the

399 

6-31G or 6-311G** basis set. The calculated absorption wavelength for the quinone dimer ranges

400 

between 348 nm and 483 nm depending on the functional and basis set used. Detailed results can

401 

be found in Table S2. We note that this calculation method does not yield information about the

402 

oscillator strength (molar absorptivity).

403 

Based on the UPLC/ESI-HR-QTOFMS mass spectra (Figure 5), UPLC/DAD (Figure 6), and the

404 

UV-Vis absorption (Figure 3), a higher concentration of the quinone dimer was in the aqueous

405 

phase than in the DCM phase, but we observed higher light absorption from the DCM phase

406 

(Figures 3, as well as Figure 6) than the filtrate (Figures 3). Therefore, we infer that additional

407 

light-absorbing species were present in the product mixture.

408 

Slikboer and coworkers 48 reported that the formation of light-absorbing products during the dark

409 

reaction of guaiacol and catechol with Fe(III) was accompanied by the formation of colloidal

ACS Paragon Plus Environment

Page 22 of 32

Page 23 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

410 

organic particles and a phase separation, similar to our observations of precipitate formation. They

411 

proposed that the phase separation was due to the formation of hydrophobic polymer products in

412 

their system. Lavi et al. also observed the formation of water-insoluble light-absorbing products

413 

from the oxidation of guaiacol, catechol and o- and p-cresol by Fe (III) under acidic conditions.49

414 

Based on our observation that surface tension stays constant after 24 hours of reaction, we infer

415 

that the reaction products are not necessarily more hydrophobic (water-insoluble) than the

416 

reactants. This is supported by the fact that light-absorbing products appeared in both aqueous and

417 

DCM extracts, as evidenced by the UV-Vis, UPLC/DAD, and UPLC/ESI-HR-QTOFMS analysis.

418 

The quinone intermediate (C16H16O6) has a similar O:C ratio to syringol (C8H10O3), 0.375. We

419 

note that aqueous inorganic/organic mixtures with similarly low O:C ratios have been observed to

420 

undergo liquid-liquid phase separation in aerosols.57-58 Phase separation initiated by the organics

421 

may lead to chemical potential-driven repartitioning of the salt and the salt/organic co-precipitation

422 

that we infer from comparing the mass of the precipitate to the initial mass of organics in the

423 

system.

424 

Atmospheric implications

425 

A previously unrecognized mechanism for the in situ formation of aerosol BrC has been identified

426 

which involves dark reactions of syringol in aqueous media containing chloride salts. Biomass

427 

burning aerosols may contain significant chloride.59 This mechanism may be especially active in

428 

air masses impacted by both biomass burning and sea salt (e.g. southern West Africa60 and the

429 

north coast of Australia61) and should be further investigated in field measurements.

430 

The formation of light-absorbing organic products was accompanied by phase separation in our

431 

experiments and those of Slikboer et al.48 Aerosol-phase experiments are necessary in order to

ACS Paragon Plus Environment

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

432 

determine if this could happen under the chemical and geometric conditions of an aerosol particle.

433 

If this does occur in atmospheric aerosols, such a phase change may have additional implications

434 

for the aerosol physicochemical properties beyond light absorption, such as aerosol heterogeneous

435 

chemistry (e.g., impeding reactive uptake of organics or trace pollutants), CCN ability, and gas-

436 

particle partitioning. BrC in atmospheric aerosols has been observed to exist as amorphous brown

437 

carbon spheres, or “tar balls”,62-65 as well as water soluble forms.66-68

438 

Acknowledgements

439 

We are grateful to Prof. Avram Gold and Prof. Zhenfa Zhang for helpful discussions. The authors

440 

acknowledge support of the National Science Foundation (CHE–1506789). Jian Xu acknowledges

441 

a scholarship fund from China Scholarship Council (File No. 201506100076) for participation in

442 

this research. Tianqu Cui and Jason D. Surratt were supported in part by NOAA (awards

443 

NA13OAR4310064 and NA16OAR4310106).

444 

Supporting Information

445 

Results of control experiments. Photos of reaction solutions, filters, filtrate, and filter extracts. UV-

446 

Vis and UPLC/ESI-HR-QTOFMS mass spectra corresponding to individual chromatophilic peaks.

447 

Specifications for purity and Fe/heavy metal content of chemicals used in the experiments. Results

448 

of DFT calculations. Method description for pendant drop tensiometry and UPLC/DAD-ESI-Q-

449 

TOFMS.

ACS Paragon Plus Environment

Page 24 of 32

Page 25 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

450  451 

References

452  453  454  455  456  457  458  459  460  461  462  463  464  465  466  467  468  469  470  471  472  473  474  475  476  477  478  479  480  481  482  483  484  485  486  487  488  489  490  491  492 

1. Gilardoni, S.; Massoli, P.; Paglione, M.; Giulianelli, L.; Carbone, C.; Rinaldi, M.; Decesari, S.; Sandrini, S.; Costabile, F.; Gobbi, G. P.; Pietrogrande, M. C.; Visentin, M.; Scotto, F.; Fuzzi, S.; Facchini, M. C., Direct observation of aqueous secondary organic aerosol from biomassburning emissions. Proc. Natl. Acad. Sci. U. S. A. 2016, 113 (36), 10013-10018. 2. Lukács, H.; Gelencsér, A.; Hammer, S.; Puxbaum, H.; Pio, C.; Legrand, M.; Kasper-Giebl, A.; Handler, M.; Limbeck, A.; Simpson, D.; Preunkert, S., Seasonal trends and possible sources of brown carbon based on 2-year aerosol measurements at six sites in Europe. J. Geophys. Res.: Atmos. 2007, 112 (D23), D23S18. 3. Saleh, R.; Hennigan, C. J.; McMeeking, G. R.; Chuang, W. K.; Robinson, E. S.; Coe, H.; Donahue, N. M.; Robinson, A. L., Absorptivity of brown carbon in fresh and photo-chemically aged biomass-burning emissions. Atmos. Chem. Phys. 2013, 13 (15), 7683-7693. 4. Andreae, M.; Gelencsér, A., Black carbon or brown carbon? The nature of light-absorbing carbonaceous aerosols. Atmos. Chem. Phys. 2006, 6 (10), 3131-3148. 5. Liu, J.; Bergin, M.; Guo, H.; King, L.; Kotra, N.; Edgerton, E.; Weber, R. J., Size-resolved measurements of brown carbon in water and methanol extracts and estimates of their contribution to ambient fine-particle light absorption. Atmos. Chem. Phys. 2013, 13 (24), 12389-12404. 6. Lack, D. A.; Langridge, J. M.; Bahreini, R.; Cappa, C. D.; Middlebrook, A. M.; Schwarz, J. P., Brown carbon and internal mixing in biomass burning particles. Proc. Natl. Acad. Sci. U. S. A. 2012, 109 (37), 14802-14807. 7. Saleh, R.; Robinson, E. S.; Tkacik, D. S.; Ahern, A. T.; Liu, S.; Aiken, A. C.; Sullivan, R. C.; Presto, A. A.; Dubey, M. K.; Yokelson, R. J.; Donahue, N. M.; Robinson, A. L., Brownness of organics in aerosols from biomass burning linked to their black carbon content. Nat. Geosci. 2014, 7 (9), 647-650. 8. Chen, Y.; Bond, T. C., Light absorption by organic carbon from wood combustion. Atmos. Chem. Phys. 2010, 10 (4), 1773-1787. 9. Budisulistiorini, S. H.; Riva, M.; Williams, M.; Chen, J.; Itoh, M.; Surratt, J. D.; Kuwata, M., Light-Absorbing Brown Carbon Aerosol Constituents from Combustion of Indonesian Peat and Biomass. Environ. Sci. Tech. 2017, 51 (8), 4415-4423. 10. Clarke, A.; McNaughton, C.; Kapustin, V.; Shinozuka, Y.; Howell, S.; Dibb, J.; Zhou, J.; Anderson, B.; Brekhovskikh, V.; Turner, H.; Pinkerton, M., Biomass burning and pollution aerosol over North America: Organic components and their influence on spectral optical properties and humidification response. J. Geophys. Res.: Atmos. 2007, 112 (D12), D12S18.

ACS Paragon Plus Environment

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

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  522  523  524  525  526  527  528  529  530  531  532  533  534  535  536  537  538 

11. Lin, P.; Aiona, P. K.; Li, Y.; Shiraiwa, M.; Laskin, J.; Nizkorodov, S. A.; Laskin, A., Molecular Characterization of Brown Carbon in Biomass Burning Aerosol Particles. Environ. Sci. Tech. 2016, 50 (21), 11815-11824. 12. Sun, H.; Biedermann, L.; Bond, T. C., Color of brown carbon: A model for ultraviolet and visible light absorption by organic carbon aerosol. Geophys. Res. Lett. 2007, 34 (17), L17813. 13. Lack, D.; Cappa, C., Impact of brown and clear carbon on light absorption enhancement, single scatter albedo and absorption wavelength dependence of black carbon. Atmos. Chem. Phys. 2010, 10 (9), 4207-4220. 14. Laing, J. R.; Jaffe, D. A.; Hee, J. R., Physical and optical properties of aged biomass burning aerosol from wildfires in Siberia and the Western USA at the Mt. Bachelor Observatory. Atmos. Chem. Phys. 2016, 16 (23), 15185-15197. 15. Laskin, A.; Laskin, J.; Nizkorodov, S. A., Chemistry of Atmospheric Brown Carbon. Chem. Rev. 2015, 115 (10), 4335-4382. 16. Yu, L.; Smith, J.; Laskin, A.; Anastasio, C.; Laskin, J.; Zhang, Q., Chemical characterization of SOA formed from aqueous-phase reactions of phenols with the triplet excited state of carbonyl and hydroxyl radical. Atmos. Chem. Phys. 2014, 14 (24), 13801-13816. 17. Lin, P.; Bluvshtein, N.; Rudich, Y.; Nizkorodov, S. A.; Laskin, J.; Laskin, A., Molecular Chemistry of Atmospheric Brown Carbon Inferred from a Nationwide Biomass Burning Event. Environ. Sci. Tech. 2017, 51 (20), 11561-11570. 18. Di Lorenzo Robert, A.; Young Cora, J., Size separation method for absorption characterization in brown carbon: Application to an aged biomass burning sample. Geophys. Res. Lett. 2015, 43 (1), 458-465. 19. Wong, J. P. S.; Nenes, A.; Weber, R. J., Changes in Light Absorptivity of Molecular Weight Separated Brown Carbon Due to Photolytic Aging. Environ. Sci. Tech. 2017, 51 (15), 8414-8421. 20. Forrister, H.; Liu, J.; Scheuer, E.; Dibb, J.; Ziemba, L.; Thornhill Kenneth, L.; Anderson, B.; Diskin, G.; Perring Anne, E.; Schwarz Joshua, P.; Campuzano‐Jost, P.; Day Douglas, A.; Palm Brett, B.; Jimenez Jose, L.; Nenes, A.; Weber Rodney, J., Evolution of brown carbon in wildfire plumes. Geophys. Res. Lett. 2015, 42 (11), 4623-4630. 21. Simoneit, B. R. T.; Rogge, W. F.; Mazurek, M. A.; Standley, L. J.; Hildemann, L. M.; Cass, G. R., Lignin pyrolysis products, lignans, and resin acids as specific tracers of plant classes in emissions from biomass combustion. Environ. Sci. Tech. 1993, 27 (12), 2533-2541. 22. Bahadur, R.; Praveen, P. S.; Xu, Y.; Ramanathan, V., Solar absorption by elemental and brown carbon determined from spectral observations. Proc. Natl. Acad. Sci. U. S. A. 2012, 109 (43), 17366-17371.

ACS Paragon Plus Environment

Page 26 of 32

Page 27 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

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  569  570  571  572  573  574  575  576  577  578  579  580  581  582  583 

ACS Earth and Space Chemistry

23. Chung, C. E.; Ramanathan, V.; Decremer, D., Observationally constrained estimates of carbonaceous aerosol radiative forcing. Proc. Natl. Acad. Sci. U. S. A. 2012, 109 (29), 1162411629. 24. McNeill, V. F., Atmospheric Aerosols: Clouds, Chemistry, and Climate. Annu. Rev. Chem. Biomol. Eng. 2017, 8 (1), 427-444. 25. Hawthorne, S. B.; Miller, D. J.; Barkley, R. M.; Krieger, M. S., Identification of methoxylated phenols as candidate tracers for atmospheric wood smoke pollution. Environ. Sci. Tech. 1988, 22 (10), 1191-1196. 26. Nolte, C. G.; Schauer, J. J.; Cass, G. R.; Simoneit, B. R. T., Highly Polar Organic Compounds Present in Wood Smoke and in the Ambient Atmosphere. Environ. Sci. Tech. 2001, 35 (10), 1912-1919. 27. Sagebiel, J. C.; Seiber, J. N., Studies on the occurrence and distribution of wood smoke marker compounds in foggy atmospheres. Environ. Toxicol. Chem. 1993, 12 (5), 813-822. 28. Romonosky, D. E.; Laskin, A.; Laskin, J.; Nizkorodov, S. A., High-Resolution Mass Spectrometry and Molecular Characterization of Aqueous Photochemistry Products of Common Types of Secondary Organic Aerosols. J. Phys. Chem. A 2015, 119 (11), 2594-2606. 29. Li, Y. J.; Huang, D. D.; Cheung, H. Y.; Lee, A. K. Y.; Chan, C. K., Aqueous-phase photochemical oxidation and direct photolysis of vanillin – a model compound of methoxy phenols from biomass burning. Atmos. Chem. Phys. 2014, 14 (6), 2871-2885. 30. Kitanovski, Z.; Čusak, A.; Grgić, I.; Claeys, M., Chemical characterization of the main products formed through aqueous-phase photonitration of guaiacol. Atmos. Meas. Tech. 2014, 7 (8), 2457-2470. 31. Chang, J. L.; Thompson, J. E., Characterization of colored products formed during irradiation of aqueous solutions containing H2O2 and phenolic compounds. Atmos. Environ. 2010, 44 (4), 541-551. 32. 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. 33. Shapiro, E. L.; Szprengiel, J.; Sareen, N.; Jen, C. N.; Giordano, M. R.; McNeill, V., Lightabsorbing secondary organic material formed by glyoxal in aqueous aerosol mimics. Atmos. Chem. Phys. 2009, 9 (7), 2289-2300. 34. Sareen, N.; Schwier, A.; Shapiro, E.; Mitroo, D.; McNeill, V., Secondary organic material formed by methylglyoxal in aqueous aerosol mimics. Atmos. Chem. Phys. 2010, 10 (3), 997-1016.

ACS Paragon Plus Environment

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

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  610  611  612  613  614  615  616  617  618  619  620  621  622  623  624  625  626  627  628  629 

35. Tang, I. N.; Munkelwitz, H. R., Water activities, densities, and refractive indices of aqueous sulfates and sodium nitrate droplets of atmospheric importance. J. Geophys. Res.: Atmos. 1994, 99 (D9), 18801-18808. 36. Tang, I. N., Thermodynamic and optical properties of mixed-salt aerosols of atmospheric importance. J. Geophys. Res.: Atmos. 1997, 102 (D2), 1883-1893. 37. Anastasio, C.; Faust, B. C.; Rao, C. J., Aromatic Carbonyl Compounds as Aqueous-Phase Photochemical Sources of Hydrogen Peroxide in Acidic Sulfate Aerosols, Fogs, and Clouds. 1. Non-Phenolic Methoxybenzaldehydes and Methoxyacetophenones with Reductants (Phenols). Environ. Sci. Tech. 1997, 31 (1), 218-232. 38. Sun, Y.; Zhang, Q.; Anastasio, C.; Sun, J., Insights into secondary organic aerosol formed via aqueous-phase reactions of phenolic compounds based on high resolution mass spectrometry. Atmos. Chem. Phys. 2010, 10 (10), 4809-4822. 39. Smith, J. D.; Sio, V.; Yu, L.; Zhang, Q.; Anastasio, C., Secondary Organic Aerosol Production from Aqueous Reactions of Atmospheric Phenols with an Organic Triplet Excited State. Environ. Sci. Tech. 2014, 48 (2), 1049-1057. 40. Zhang, H.; Surratt, J. D.; Lin, Y. H.; Bapat, J.; Kamens, R. M., Effect of relative humidity on SOA formation from isoprene/NO photooxidation: enhancement of 2-methylglyceric acid and its corresponding oligoesters under dry conditions. Atmos. Chem. Phys. 2011, 11 (13), 6411-6424. 41. Riva, M.; Budisulistiorini, S. H.; Chen, Y.; Zhang, Z.; D’Ambro, E. L.; Zhang, X.; Gold, A.; Turpin, B. J.; Thornton, J. A.; Canagaratna, M. R.; Surratt, J. D., Chemical Characterization of Secondary Organic Aerosol from Oxidation of Isoprene Hydroxyhydroperoxides. Environ. Sci. Tech. 2016, 50 (18), 9889-9899. 42. Zhao, Y.; Truhlar, D. G., The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120 (1), 215-241. 43. Surratt, J. D.; Murphy, S. M.; Kroll, J. H.; Ng, N. L.; Hildebrandt, L.; Sorooshian, A.; Szmigielski, R.; Vermeylen, R.; Maenhaut, W.; Claeys, M.; Flagan, R. C.; Seinfeld, J. H., Chemical Composition of Secondary Organic Aerosol Formed from the Photooxidation of Isoprene. J. Phys. Chem. A 2006, 110 (31), 9665-9690. 44. Kalberer, M.; Sax, M.; Samburova, V., Molecular Size Evolution of Oligomers in Organic Aerosols Collected in Urban Atmospheres and Generated in a Smog Chamber. Environ. Sci. Tech. 2006, 40 (19), 5917-5922. 45. Karas, M.; Ehring, H.; Nordhoff, E.; Stahl, B.; Strupat, K.; Hillenkamp, F.; Grehl, M.; Krebs, B., Matrix-assisted laser desorption/ionization mass spectrometry with additives to 2,5dihydroxybenzoic acid. Org. Mass Spectrom. 1993, 28 (12), 1476-1481.

ACS Paragon Plus Environment

Page 28 of 32

Page 29 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

630  631  632  633  634  635  636  637  638  639  640  641  642  643  644  645  646  647  648  649  650  651  652  653  654  655  656  657  658  659  660  661  662  663  664  665  666  667  668  669  670  671  672  673 

ACS Earth and Space Chemistry

46. Schmalzl, K. J.; Forsyth, C. M.; Evans, P. D., The reaction of guaiacol with iron III and chromium VI compounds as a model for wood surface modification. Wood Sci. Technol. 1995, 29 (4), 307-319. 47. Hwang, S.; Lee, C.-H.; Ahn, I.-S., Product identification of guaiacol oxidation catalyzed by manganese peroxidase. J. Ind. Eng. Chem. 2008, 14 (4), 487-492. 48. Slikboer, S.; Grandy, L.; Blair, S. L.; Nizkorodov, S. A.; Smith, R. W.; Al-Abadleh, H. A., Formation of Light Absorbing Soluble Secondary Organics and Insoluble Polymeric Particles from the Dark Reaction of Catechol and Guaiacol with Fe(III). Environ. Sci. Tech. 2015, 49 (13), 77937801. 49. Lavi, A.; Lin, P.; Bhaduri, B.; Carmieli, R.; Laskin, A.; Rudich, Y., Characterization of Light-Absorbing Oligomers from Reactions of Phenolic Compounds and Fe(III). ACS Earth Space Chem. 2017, 1 (10), 637-646. 50. MacArthur, C. G., Solubility of Oxygen in Salt Solutions and the Hydrates of These Salts. J. Phys. Chem. 1915, 20 (6), 495-502. 51. Burant, A.; Lowry, G. V.; Karamalidis, A. K., Measurement and Modeling of Setschenow Constants for Selected Hydrophilic Compounds in NaCl and CaCl2 Simulated Carbon Storage Brines. Acc. Chem. Res. 2017, 50 (6), 1332-1341. 52. Setschenow, J., Über die konstitution der salzlösungen auf grund ihres verhaltens zu kohlensäure. Z. Phys. Chem. 1889, 4 (1), 117-125. 53. Israelachvili, J., Intermolecular and Surface Forces: Revised Third Edition. Intermolecular and Surface Forces. Elsevier Science: 2011. 54. Toivola, M.; Prisle, N. L.; Elm, J.; Waxman, E. M.; Volkamer, R.; Kurtén, T., Can COSMOTherm Predict a Salting in Effect? J. Phys. Chem. A 2017, 121 (33), 6288-6295. 55. Zakzeski, J.; Bruijnincx, P. C. A.; Weckhuysen, B. M., In situ spectroscopic investigation of the cobalt-catalyzed oxidation of lignin model compounds in ionic liquids. Green Chem. 2011, 13 (3), 671-680. 56. Zakzeski, J.; Jongerius, A. L.; Weckhuysen, B. M., Transition metal catalyzed oxidation of Alcell lignin, soda lignin, and lignin model compounds in ionic liquids. Green Chem. 2010, 12 (7), 1225-1236. 57. You, Y.; Smith, M. L.; Song, M.; Martin, S. T.; Bertram, A. K., Liquid–liquid phase separation in atmospherically relevant particles consisting of organic species and inorganic salts. Int. Rev. Phys. Chem. 2014, 33 (1), 43-77.

ACS Paragon Plus Environment

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

674  675  676  677  678  679  680  681  682  683  684  685  686  687  688  689  690  691  692  693  694  695  696  697  698  699  700  701  702  703  704  705  706  707  708  709  710  711  712  713  714  715  716  717  718  719 

58. Song, M.; Marcolli, C.; Krieger, U. K.; Zuend, A.; Peter, T., Liquid-liquid phase separation in aerosol particles: Dependence on O:C, organic functionalities, and compositional complexity. Geophys. Res. Lett. 2012, 39 (19), L19801. 59. Ahern, A. T.; Goldberger, L.; Jahl, L.; Thornton, J.; Sullivan, R. C., Production of N2O5 and ClNO2 through Nocturnal Processing of Biomass-Burning Aerosol. Environ. Sci. Tech. 2018, 52 (2), 550-559. 60. Groß, S.; Tesche, M.; Freudenthaler, V.; Toledano, C.; Wiegner, M.; Ansmann, A.; Althausen, D.; Seefeldner, M., Characterization of Saharan dust, marine aerosols and mixtures of biomass-burning aerosols and dust by means of multi-wavelength depolarization and Raman lidar measurements during SAMUM 2. Tellus B 2011, 63 (4), 706-724. 61. Bouya, Z.; Box, G. P.; Box, M. A., Seasonal variability of aerosol optical properties in Darwin, Australia. J. Atmos. Sol.-Terr. Phys. 2010, 72 (9), 726-739. 62. Meskhidze, N.; Petters, M. D.; Tsigaridis, K.; Bates, T.; O'Dowd, C.; Reid, J.; Lewis, E. R.; Gantt, B.; Anguelova, M. D.; Bhave, P. V.; Bird, J.; Callaghan, A. H.; Ceburnis, D.; Chang, R.; Clarke, A.; de Leeuw, G.; Deane, G.; DeMott, P. J.; Elliot, S.; Facchini, M. C.; Fairall, C. W.; Hawkins, L.; Hu, Y. X.; Hudson, J. G.; Johnson, M. S.; Kaku, K. C.; Keene, W. C.; Kieber, D. J.; Long, M. S.; Martensson, M.; Modini, R. L.; Osburn, C. L.; Prather, K. A.; Pszenny, A.; Rinaldi, M.; Russell, L. M.; Salter, M.; Sayer, A. M.; Smirnov, A.; Suda, S. R.; Toth, T. D.; Worsnop, D. R.; Wozniak, A.; Zorn, S. R., Production mechanisms, number concentration, size distribution, chemical composition, and optical properties of sea spray aerosols. Atmos. Sci. Lett. 2013, 14 (4), 207-213. 63. Posfai, M.; Gelencser, A.; Simonics, R.; Arato, K.; Li, J.; Hobbs, P. V.; Buseck, P. R., Atmospheric tar balls: Particles from biomass and biofuel burning. J. Geophys. Res.: Atmos. 2004, 109 (D6), D06213. 64. Chakrabarty, R. K.; Moosmüller, 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. Alexander, D. T.; Crozier, P. A.; Anderson, J. R., Brown carbon spheres in East Asian outflow and their optical properties. Science 2008, 321 (5890), 833-836. 66. Hoffer, A.; Gelencsér, A.; Guyon, P.; Kiss, G.; Schmid, O.; Frank, G. P.; Artaxo, P.; Andreae, M. O., Optical properties of humic-like substances (HULIS) in biomass-burning aerosols. Atmos. Chem. Phys. 2006, 6 (11), 3563-3570. 67. Graber, E. R.; Rudich, Y., Atmospheric HULIS: How humic-like are they? A comprehensive and critical review. Atmos. Chem. Phys. 2006, 6 (3), 729-753. 68. Liu, J.; Scheuer, E.; Dibb, J.; Diskin, G. S.; Ziemba, L. D.; Thornhill, K. L.; Anderson, B. E.; Wisthaler, A.; Mikoviny, T.; Devi, J. J.; Bergin, M.; Perring, A. E.; Markovic, M. Z.; Schwarz,

ACS Paragon Plus Environment

Page 30 of 32

Page 31 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

720  721  722 

ACS Earth and Space Chemistry

J. P.; Campuzano-Jost, P.; Day, D. A.; Jimenez, J. L.; Weber, R. J., Brown carbon aerosol in the North American continental troposphere: sources, abundance, and radiative forcing. Atmos. Chem. Phys. 2015, 15 (14), 7841-7858.

723 

ACS Paragon Plus Environment

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

724  725 

FOR TOC ONLY

726 

ACS Paragon Plus Environment

Page 32 of 32