Maillard Chemistry in Clouds and Aqueous Aerosol As a Source of

May 26, 2016 - For example, browning occurs when limonene SOA is exposed to NH3(g) or NH4+(aq)(13) and when small aldehydes undergo aldol condensation...
0 downloads 12 Views 2MB Size
Subscriber access provided by The University of Melbourne Libraries

Article

Maillard chemistry in clouds and aqueous aerosol as a source of atmospheric humic-like substances. Lelia Nahid Hawkins, Amanda N. Lemire, Melissa Marie Galloway, Ashley L. Corrigan, Jacob J. Turley, Brenna M. Espelien, and David O De Haan Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b00909 • Publication Date (Web): 26 May 2016 Downloaded from http://pubs.acs.org on June 5, 2016

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

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

Page 1 of 31

Environmental Science & Technology

1

Maillard chemistry in clouds and aqueous aerosol as

2

a source of atmospheric humic-like substances.

3

Lelia N. Hawkins,a,b,* Amanda N. Lemire,b Melissa M. Galloway,a,c Ashley L. Corrigan,a Jacob J.

4

Turley,a Brenna M. Espelien,a David O. De Haana

5

a. Department of Chemistry and Biochemistry, University of San Diego, 5998 Alcala Park, San

6

Diego CA 92110

7

b. Currently at Department of Chemistry, Harvey Mudd College, 301 Platt Blvd., Claremont CA

8

91711

9

c. Currently at Department of Chemistry, Lafayette College, 730 High St, Easton, PA 18042

10 11

*Corresponding Author:

12

Lelia Hawkins: Harvey Mudd College, 301 Platt Blvd., Claremont, CA 91711-5990, Phone:

13

909.621.8522, Fax: 909.607.7577, [email protected].

14

KEYWORDS: secondary organic aerosol formation, oligomers, light-absorbing aerosol, cloud

15

processing, STXM.

16

ACS Paragon Plus Environment

1

Environmental Science & Technology

Page 2 of 31

17

ABSTRACT. The reported optical, physical, and chemical properties of aqueous Maillard

18

reaction mixtures of small aldehydes (glyoxal, methylglyoxal, and glycolaldehyde) with

19

ammonium sulfate and amines are compared with those of aqueous extracts of ambient aerosol

20

(water-soluble organic carbon, WSOC) and the humic-like substances (HULIS) fraction of

21

WSOC. Using a combination of new and previously published measurements, we examine

22

fluorescence, X-ray absorbance, UV/vis, and IR spectra, complex refractive indices, 1H and 13C

23

NMR spectra, thermograms, aerosol and electrospray ionization mass spectra, surface activity,

24

and hygroscopicity. Atmospheric WSOC and HULIS encompass a range of properties, but in

25

almost every case, aqueous aldehyde-amine reaction mixtures are squarely within this range.

26

Notable exceptions are the higher UV-visible absorbance wavelength dependence (Angstrom

27

coefficients) observed for methylglyoxal reaction mixtures, the lack of surface activity of glyoxal

28

reaction mixtures, and the higher N/C ratios of aldehyde-amine reaction products relative to

29

atmospheric WSOC and HULIS extracts. The overall optical, physical, and chemical similarities

30

are consistent with, but not demonstrative of, Maillard chemistry being a significant secondary

31

source of atmospheric HULIS. However, the higher N/C ratios of aldehyde-amine reaction

32

products limits the source strength to ≤ 50% of atmospheric HULIS, assuming that other sources

33

of HULIS incorporate only negligible quantities of nitrogen.

34 35 36 37 38 39

ACS Paragon Plus Environment

2

Page 3 of 31

Environmental Science & Technology

40

Introduction

41

In spite of significant efforts, the molecular components of atmospheric organic aerosol remain

42

largely unidentified. A primary difficulty is the presence of thousands of different compounds,

43

some of which react in aqueous or particulate-phase reactions to form complex oligomers. For

44

example, single particle measurements in late summer and fall of 2005 indicate that half of

45

organic aerosol particles in Los Angeles were significantly oligomerized (1), and atmospheric

46

oligomers have been found to contain organic nitrogen (2). Aerosol oligomers are largely water-

47

soluble, associated with light-absorbing organic compounds (brown carbon, BrC), and similar to

48

terrestrial humic and fulvic acids, resulting in the acronym HULIS (humic-like substances) (3) as

49

a collective term for oligomerized aerosol material. It is now clear, however, that lofted

50

terrestrial humic and fulvic acids are not a major source of oligomerized aerosol or BrC, since

51

terrestrial acids have larger molecular weight, higher aromaticity, lower surface activity, poorer

52

droplet activation ability, and longer fluorescence wavelengths than typical aerosol extracts (4).

53

Additionally, growing evidence suggests a secondary, HULIS BrC production pathway exists in

54

clouds and aerosol (4-6). The production of light-absorbing and oligomeric compounds has

55

important implications for climate (7-9), aerosol-cloud interactions (10), and the rate of diffusion

56

of water and reactive gases into aerosol particles due to increasing viscosity and associated phase

57

changes (11). A comprehensive review of current measurements and models of brown carbon

58

(including HULIS BrC) composition, reactivity, and optical properties can be found in Laskin

59

2015 (12). A better understanding of the sources and properties of HULIS BrC is needed to

60

improve predictions of climate change and particulate air pollution.

61

Several aqueous and particle-phase reactions have been suggested as sources of brown and/or

62

oligomerized material in the atmosphere. For example, browning occurs when limonene SOA is

ACS Paragon Plus Environment

3

Environmental Science & Technology

Page 4 of 31

63

exposed to NH3(g) or NH4+(aq) (13) and when small aldehydes undergo aldol condensation

64

catalyzed by amino acids and ammonium ions (14-16).

65

efficient for producing BrC in solutions containing glycolaldehyde and methylglyoxal (17). OH

66

radical oxidation of phenols quickly produces brown products (18), and aldehyde-radical

67

reactions produce oligomers under aerosol-like conditions (19-21). Glyoxal, methylglyoxal,

68

glycolaldehyde, or hydroxyacetone react with ammonium sulfate (AS) (17, 22-27) or primary

69

amines (17, 28, 29) to produce imine oligomers, N-containing aromatic derivatives, and a small

70

amount of highly absorbing brown products. The lifetime of both brown and high molecular

71

weight products from these reactions may be short during daytime in both clean and polluted air

72

masses due to photolysis (30), though more studies are needed to determine if all AS-aldehyde

73

systems and any amine-aldehyde systems behave as methylglyoxal-AS. These latter processes

74

are key elements of Maillard reactions (sugars + proteins), since sugars and proteins break down

75

upon heating into small aldehydes and individual amino acids. While aldehyde-amine reactions

76

are slow in bulk aqueous-phase studies at room temperature, they are accelerated in drying

77

aerosol particles by a factor of ~103 beyond any concentration effect (28), increasing the mass of

78

the residual dried aerosol (31). Although future studies are needed to determine the acceleration

79

of these reactions under intermediate relative humidity conditions, this acceleration makes

80

Maillard reactions potentially competitive with OH radical reactions as sinks for small aldehydes

81

in non-acidified aerosol (32).

Aldol condensation is especially

82

Here we analyze the products of bulk aqueous-phase reactions of the common atmospheric

83

aldehydes glycolaldehyde (GAld), glyoxal (GX), and methyglyoxal (MG) with AS and amines.

84

Rather than study the kinetics of formation, our goal is to compare the reported characteristics of

85

these products with published measurements of reference humic substances and atmospheric

ACS Paragon Plus Environment

4

Page 5 of 31

Environmental Science & Technology

86

particulates (including extracts) to determine whether these reactions are a plausible atmospheric

87

source of secondary HULIS-type BrC. Using a combination of new and previously published

88

measurements, we find that Maillard reaction mixtures of atmospheric aldehydes and amines

89

match a remarkable array of properties observed for atmospheric HULIS BrC, with very few

90

exceptions.

91 92

Materials and Methods

93

Sample preparation procedures for all new measurements are described below. All chemicals

94

were used as received from Sigma-Aldrich unless otherwise designated.

95

Fluorescence. 1M stock solutions were made by hydrolyzing para-formaldehyde polymer

96

(95%) or glyoxal trimer dihydrate (>95%, Fluka) in 18 MΩ deionized water. Each reaction

97

mixture contained 0.25 M of an aldehyde and either glycine (>99%) or ammonium sulfate

98

(>99%), was buffered to pH 4 with acetic acid, and reacted at room temperature. Excitation –

99

emission fluorescence maps from 220 nm were collected twice per week in 1×1 cm quartz

100

cuvettes (JASCO FP-6500, 0.01 s response, high sensitivity, 200 nm/min emission scan speed, 5

101

nm data pitch, 3 nm spectral bandwidths).

102

FTIR. Solutions contained 0.05 M of an aldehyde freshly mixed with 0.1 M of an amino acid

103

in 18 MΩ deionized water, and had pH = 5.3 to 6.0. 1 µL aliquots were dried for 10 min. under

104

ambient conditions on attenuated total reflectance (ATR) 9-reflection diamond crystals

105

(DuraSamplIR, Smiths Detection) and analyzed by Fourier Transform Infrared (FTIR)

106

spectroscopy (JASCO 4800, 0.5 cm-1 resolution).

107

Nuclear Magnetic Resonance and Electrospray Ionization-MS. Aqueous samples containing

108

0.5 M methylglyoxal (diluted from 40% aqueous solution, Alfa-Aesar), glyoxal, or

ACS Paragon Plus Environment

5

Environmental Science & Technology

Page 6 of 31

109

glycolaldehyde (hydrolyzed from dimer) and 0.5 M amine (methylamine, diluted from 40%

110

aqueous solution; glycine, serine, >99%; arginine, >98%; or ornithine-HCl, 99%) were mixed in

111

glass vials and evaporated under ambient conditions. Dried residues were redissolved in D2O

112

(Varian Inova 500 MHz, 5000 scan 13C or 8 scan 1H spectra recorded) or 18 MΩ deionized water

113

at 1 mg/mL (Thermo Finnigan LCQ ESI-MS, +4 kV, 250 °C capillary, syringe-pumped at 1

114

µL/min, sheath flow at 20).

115

STXM-NEXAFS. Aqueous samples were prepared and dried following the NMR protocol.

116

The dried residues were redissolved in 18 MΩ deionized water, atomized using a Brechtel BMI

117

Aerosol Generation System, collected by impaction on silicon nitride slides (Si3N4, Silson, Ltd.,

118

Northampton, England), and frozen until spectroscopic analysis. Scanning Transmission X-ray

119

Microscopy – Near Edge X-ray Absorption Fine Structure (STXM-NEXAFS) has previously

120

been employed to characterize individual organic aerosol particles at the carbon K-edge (33-36).

121

Particles were analyzed on Beamline 5.3.2 at the Lawrence Berkeley National Laboratory

122

Advanced Light Source facility. Details of the energy calibration, sample handling, and data

123

analysis are presented in (34). Each particle is scanned from 280 to 320 eV; the scan results in a

124

pixelated image of each particle based on carbon absorption intensity at any point along the

125

carbon K-edge. A post-analysis Matlab (Mathworks) processing algorithm was used to define the

126

particle in each stack of images and obtain particle-average spectra (33).

127 128 129

Results and Discussion Fluorescence.

Many atmospheric aerosol particles, especially in biogenic-dominated air

130

masses, are strongly fluorescent.

The fluorescence of atmospheric HULIS BrC was

131

characterized by Graber and Rudich (4) with excitation / emission peaks at 340/450 nm (“fulvic-

ACS Paragon Plus Environment

6

Page 7 of 31

Environmental Science & Technology

132

like”) and 250/430 nm (“humic-like”). SOA produced in the α-pinene + O3 system becomes

133

strongly fluorescent in the presence of NH3 (13). Aqueous-phase reactions of small aldehydes

134

with AS and amines also generate highly fluorescent products (17). The fluorescence of a few of

135

these reaction systems is compared with atmospheric WSOC collected in Portugal (37) in Figure

136

1. Glyoxal-AS and glyoxal-glycine reaction products fluoresce at wavelengths that match the

137

fulvic-like fluorescence peak commonly seen in atmospheric HULIS BrC, while formaldehyde-

138

AS reaction product fluorescence matches the humic-like peak.

139

fluorescence wavelengths of the products of these three reactions are stable even at long reaction

140

times in the bulk phase (17). Other aldehydes, such as MG and GAld, also generate intensely

141

fluorescent product mixtures when reacting with AS and amines, but peaks shift to longer

142

wavelengths during the course of reaction due to lengthening of conjugated products by aldol

143

condensation (17).

Additionally, the peak

144 145

UV-Vis. Many BrC candidate aqueous-phase reactions studied thus far in the lab generate

146

product mixtures exhibiting discrete absorption bands in the UV or visible region, such as

147

limonene SOA + AS (13, 38), syringol + OH (18), or acetaldehyde + amino acids (15). These

148

“banded” spectra contrast with the featureless continuum usually observed in BrC extracted from

149

atmospheric aerosol (39).

150

absorption bands could produce a mixture exhibiting a smooth absorbance continuum, even if

151

many such reactions are happening concurrently (40, 41). Instead, such featureless absorption

152

and fluorescence spectra have been attributed to the formation of a complex mixture of products

153

with carbonyl and hydroxyl functional groups, energetically linked to function as charge transfer

154

complexes (42).

It is unlikely that reactions producing compounds with discrete

ACS Paragon Plus Environment

7

Environmental Science & Technology

Page 8 of 31

155

Continuum absorbance spectra are well-fitted by a power law with exponent -λ, where λ is the

156

Angstrom coefficient (4). While λ ~ 1 for black carbon (12), absorption spectra of BrC are

157

characterized by steeper wavelengths dependence with 4 < λ < 8 (5, 43, 44). This type of data

158

produces a straight line when graphed on a log scale (Figure 2) with slopes that correlate with –λ.

159

Thus, one way such “featureless” chemical systems might be distinguished is by these slopes. In

160

Figure 2, published aldehyde-AS and aldehyde-amine absorption data (17) is compared with

161

urban WSOC (5). Linear fits are shown for each aldehyde. The wavelength dependence (or

162

slope) of the absorbance of GX-amine and GX-amine-AS reactions (shown in red) between 300

163

and 600 nm matches perfectly that observed in aqueous extracts of urban and biomass-burning-

164

influenced aerosol (shown in green) (5), while MG-amine reactions (shown in black) have a

165

significantly steeper wavelength dependence. GAld-amine reactions (shown in blue) produce

166

~10× more light absorption per mole than glyoxal reactions, yet the average slope is still close to

167

that observed in atmospheric HULIS BrC. This analysis suggests that it is at least plausible that

168

glyoxal- and glycolaldehyde-amine reactions are major contributors to the atmospheric BrC

169

sampled by Hecobian et al. (5), but that methylglyoxal-amine reactions are not dominant sources

170

of BrC in these aerosol samples.

171 172

Refractive index. Complex refractive indices were determined previously by cavity ringdown

173

spectroscopy at 532 nm on products of GX and MG + amine reactions conducted without pH

174

control (45). Here we compare those values to reported refractive indices for atmospheric

175

samples (at 532-550 nm). For unbuffered reactions with glycine and methylamine, pH ranges

176

from 3 - 5 and 8 – 10, respectively. With aldehyde-glycine reactions, acidity comes from the

177

pyruvic acid impurity in MG and from products such as formic acid (22).

The complex

ACS Paragon Plus Environment

8

Page 9 of 31

Environmental Science & Technology

178

refractive indices for GX-amine and MG-glycine aerosol (Figure 3) match that of atmospheric

179

HULIS BrC extracted from smoke and from polluted environments, while aerosol particles

180

generated from MG-methylamine reaction products (at high pH) were slightly more absorbing

181

and less scattering.

182

collected in the Amazon, and much less absorbing than BrC spheres (“tar balls”) or black carbon

183

/ soot. From these measurements we conclude that aldehyde-amine aqueous BrC materials

184

appear to be optically indistinguishable from BrC associated with biomass burning and urban

185

BrC.

The aldehyde-amine aerosol were much more absorbing than HULIS

186 187

FTIR. The FTIR spectra of four aqueous droplets containing an aldehyde and an amino acid,

188

dried on the ATR crystal surface, are shown in Figure 4, along with those of urban aerosol

189

HULIS extracts (39, 46).

190

simultaneously in evaporating cloud droplets, the average of these spectra is also shown for

191

comparison. While similarities in FTIR spectra of complex mixtures do not necessarily imply

192

similar structures (4), spectral differences do imply significant differences in the predominant

193

functional groups present. It is notable that the Guangzhou HULIS extract has a stronger

194

shoulder at 3550 cm-1 due OH or NH stretching, while the Copenhagen HULIS extract matches

195

the reaction samples more closely in this range. The glyoxal – amine samples are similar to the

196

Copenhagen HULIS extract in the 2450 – 2900 cm-1 range where carboxylate OH and protonated

197

amine NH stretches are dominant. Both HULIS extracts and 3 out of 4 aldehyde – amine reaction

198

mixtures exhibit a peak (or shoulder) at 1720 cm-1, likely due to protonated carboxylic acid

199

functional groups. Additional discussion of peak assignments can be found in the SI. Overall, the

Since many aldehyde – amino acid reactions could take place

ACS Paragon Plus Environment

9

Environmental Science & Technology

Page 10 of 31

200

aldehyde – amino acid products have significantly higher C-N and C-O signals and slightly

201

lower O-H signals than the HULIS extracts.

202 203

NMR. A 1H-NMR spectrum of a dried MG – methylamine reaction mixture redissolved in

204

D2O is shown in Figure 5, along with atmospheric organic aerosol spectra taken from the

205

literature ((3, 47). In the MG – methylamine reaction mixture, signals assigned to unreacted MG

206

(1.39 and 5.25 ppm) (32), unreacted methylamine (2.59 ppm), and bulk-phase product 1,3,4-

207

trimethylimidazole (7.16 and 8.53 ppm) (32), are not observed. Instead, prominent narrow peaks

208

are seen at 0.96, 1.75, 2.03, 2.42, and 8.27 ppm along with broad groups of unresolved signals

209

centered at 1.25, 2.6, and 3.6 ppm that indicate a very complex mixture of products.

210

Interestingly, the narrow 2.42 ppm peak is the strongest (non-solvent) signal in both the MG-

211

methylamine reaction products spectrum and the HULIS extract from NIST 1648 urban dust (3).

212

A broader peak at 1.25 ppm is also present in both spectra. While the large number of aerosol

213

NMR spectra used to generate the PMF factor spectrum eliminate most narrow NMR peaks due

214

to individual compounds, the resulting three broad peaks labeled “HC- O” “HC-C=O” and “HC”

215

in the HULIS factor (Figure 5c) match the location of broad peaks in the MG-methylamine

216

reaction mixture more closely than the other PMF factor spectra do. The relative sizes of the

217

broad peaks are different, however, with the atmospheric HULIS spectra containing greater

218

signals from aliphatic hydrogens (“HC”).

219

In Figure 6 we compare an overlay of 13C NMR spectra for aldehyde-amine reaction mixtures 13

220

with rural HULIS extracts (48). The strongest

C peak observed in the HULIS extract is the

221

unsubstituted aliphatic carbon peak centered at 30 ppm, which matches the location of aliphatic

222

C in arginine and ornithine side chains in the reaction mixtures. Signals due to carbon atoms

ACS Paragon Plus Environment

10

Page 11 of 31

Environmental Science & Technology

223

singly bound to an N or O atom appear between 40 and 65 ppm. Most of these peaks might

224

contribute to the broad 30 and 55 ppm peaks observed in the HULIS extract, but not the broad

225

peak at 70 ppm, which we assign to C singly bound to both N and O. Only aldehyde-arginine

226

products have similar shifts, due to carbons labeled A and B in Figure 7 inset (49). The strong

227

peak at 90.5 ppm is observed only in GX reaction mixtures, and is due to the dissolved and fully

228

hydrated glyoxal monomer. In solid aerosol extracts, this peak would be replaced by glyoxal

229

acetal oligomer peaks in the range 90 – 105 ppm, caused by carbons singly bonded to two

230

oxygen atoms. Aromatic C peaks from 110 – 160 ppm in the HULIS extract were more

231

prominent in autumn samples influenced by biomass burning, likely due to the presence of lignin

232

breakdown products which also enhanced carbonyl signals near 200 ppm. Aromatic signals in

233

aldehyde – amine reaction mixtures are due to the formation of imidazole derivatives, with C=C

234

and C=N carbons appearing near 123 and 137 ppm, respectively. No carbonyl signals are

235

observed near 200 ppm in aldehyde – amine reaction mixtures because of favorable hydration of

236

aldehyde functional groups in these aqueous samples. Both aldehyde – amine reaction mixtures

237

and the rural HULIS extracts exhibit a carboxylic acid peak centered at 175 ppm. To summarize,

238

the most notable difference in 13C NMR signals between the HULIS extract and the aldehyde +

239

amino acid reaction mixtures is the large peak observed in the HULIS extract at 70 ppm,

240

assigned to aliphatic carbons bonded to O and N. Very few signals are observed in this region in

241

aldehyde + amino acid reaction mixtures.

242 243

AMS. Factor analysis of aerosol mass spectrometry (AMS) field spectra has been used to

244

identify a typical HULIS mass spectrum associated with continental secondary organic aerosol

245

production (47).

The predominant fragments in the HULIS AMS spectrum are m/z = 44

ACS Paragon Plus Environment

11

Environmental Science & Technology

Page 12 of 31

246

(carboxylate group) and other oxygenated or heteroatom fragments (m/z = 17, 18, 29, and 40),

247

but it was determined that polycarboxylic acids, which compose a significant fraction of HULIS

248

extracts, made up only 27% of this type of aerosol material. De Haan et al. (29) reported N/C

249

and O/C ratios of several aldehyde + amine aerosols using HR-ToF-AMS. These values are

250

compared in Figure S3 to elemental analysis results of rural and urban HULIS extracts (50-52).

251

HULIS extracts have O/C ratios (0.45 – 0.58) that are consistent with those observed in freshly

252

generated aerosol containing only MG, MG-AS, or any MG-amine mixture (except ornithine).

253

N/C ratios in atmospheric HULIS extracts (0.04 – 0.05) are at least a factor of 2 lower than those

254

of any aerosol generated from 1:1 mixtures of MG and AS or amine. The lower N/C ratios of

255

HULIS could be due to aldehyde-amine reactions occurring under conditions where

256

aldehyde:amine ratios are above 1:1, or to mixing of aldehyde-amine reaction products with non-

257

nitrogen-containing material.

258

methylglyoxal-amine reaction products and HULIS extracts can be used to place an upper limit

259

on the contribution of the reactions of C1 – C3 carbonyl compounds with amines and AS to the

260

atmospheric formation of HULIS. Using the lowest N/C ratio observed in lab experiments (MG

261

+ serine bulk reaction, N/C = 0.091) and the higher of the HULIS average N/C ratios (0.045), an

262

upper limit of ≤ 50% is calculated, which assumes that other HULIS-producing reactions do not

263

significantly incorporate nitrogen. We note that many other proposed sources of HULIS, such as

264

acid-catalyzed terpenoid or IEPOX reactions (53) or OH oxidation of phenols (18) or aldehydes

265

(19-21), do not generate nitrogen-containing products. Additionally, the consistency of O/C

266

ratios suggests that the “aging” influence of aldehyde-amine and aldehyde-AS reactions on the

267

elemental composition of atmospheric HULIS BrC will be to increase the N/C ratio while having

268

little effect on the O/C ratio.

In the latter case, the difference in N/C ratios between

ACS Paragon Plus Environment

12

Page 13 of 31

Environmental Science & Technology

269 270

ESI-MS. The clearest evidence for oligomerization in HULIS extracts of atmospheric aerosol

271

has come from electrospray ionization (ESI-) MS studies. Oligomers measured in negative ion

272

mode show a broad distribution of masses centered around m/z 300 – 350, with series of peaks 2,

273

14, and 16 amu apart (39, 54, 55). Most aldehyde + amine mixtures produce complex mixtures

274

of oligomers with similar mass patterns (28, 29), although higher N/C ratios result in better

275

detection in the positive ion mode. Figure 7 compares three HULIS extracts (negative mode)

276

(39) with a dried and redissolved glycolaldehyde + methylamine mixture (positive mode). The

277

lab generated sample shows a similar complexity and has a mass distribution centered at 330 ±6,

278

in between those of the HULIS extracts. Laskin et al. (56) used high-resolution positive ion

279

mode ESI-MS to identify N-heterocyclic alkaloid compounds as a substantial fraction of

280

nitrogen-containing organic compounds in biomass-burning aerosol samples.

281

abundant homologous series of oligomers were methylated imidazoles and pyrazines, which are

282

also major aqueous-phase products of dicarbonyls and hydroxyaldehydes reacting with AS and

283

amines (22, 25, 49). While the study notes that alkaloid compounds like these are abundant in

284

pine foliage, these compounds could also be produced by the Maillard reaction of carbohydrates

285

in the presence of proteins at elevated temperatures, conditions where BrC would also be

286

produced. Since N-heterocyclic alkaloid compounds are produced by both biomass burning and

287

aqueous aldehyde-amine Maillard chemistry, these compounds are not usable as tracers for only

288

one of these BrC sources.

The most

289 290

STXM-NEXAFS. Carbon K-edge X-ray absorbance spectra from GX-methylamine reaction

291

mixtures share more features with atmospheric HULIS samples than with humic reference

ACS Paragon Plus Environment

13

Environmental Science & Technology

Page 14 of 31

292

material (Figure 8). Based on relative peak area, it is clear that fulvic acid fractions have fewer

293

and narrower peaks and a greater aromatic content than atmospheric samples, consistent with

294

previous studies (4). The atmospheric samples shown were all influenced by biomass burning

295

(36, 57). In the region between 288.5 and 290 eV containing both O-substituted alkyl carbon (36)

296

and C-N carbon absorbances (58), fulvic acid standards have little absorbance while atmospheric

297

samples have significant absorbance.

298

absorbance at 285.1 eV than either reference material or atmospheric samples, indicating less

299

aromatic C=C content. However, the atmospheric HULIS sample containing slightly aged smoke

300

shares the broad, featureless absorbance between 284 and 287 eV. The absorbance between 286

301

and 287 eV can be attributed to C=N bonds and O-substituted aromatic carbon, among other

302

groups, making this region typically too ambiguous for specific group assignment. The GX-

303

methylamine reaction mixture product particles appear to have less carboxyl carbonyl and more

304

amide carbonyl signature, with the large peak near 288 eV shifted to lower energy than in the

305

atmospheric samples. This is consistent with elemental analyses showing the higher N/C ratio of

306

aldehyde-amine reaction products relative to atmospheric samples and with the proposed

307

products (28). The overall similarity between the GX-methylamine product spectrum and the

308

atmospheric samples supports the notion that products formed from small aldehydes and amines

309

are chemically similar to many types of atmospheric HULIS.

Glyoxal-methylamine reaction mixtures have less

310 311

HTDMA. Water uptake measurements of HULIS show continuous uptake without distinct

312

deliquescence, efflorescence, or hysteresis (59, 60). This type of behavior is observed for

313

complex mixtures. The reactions of GX, MG, and GAld with amines produce complex mixtures

314

with very similar continuous water uptake behavior (61). A useful parameter, κ, was introduced

ACS Paragon Plus Environment

14

Page 15 of 31

Environmental Science & Technology

315

by Petters and Kreidenweis (62) in order to describe hygroscopicity of dried aerosol using a

316

single term. To determine if these Maillard reactions produce material with similar

317

hygroscopicity to atmospheric HULIS, hygroscopicity kappa parameters (κ) are calculated from

318

measured growth factors for all of these systems reported in (61). The results of this comparison

319

are shown in Figure S1. Atmospheric HULIS extracts show a wide range of measured kappa

320

parameters, from 0.02 (winter rural forest) to 0.26 (polluted urban) (59). This range includes

321

kappa parameters measured for GX, MG, and GAld-amine reaction products (κ = 0.05 to 0.2),

322

but also the hygroscopicity of many other organic species, including ozonolysis SOA generated

323

from a variety of precursors (63).

324 325

Thermal profiling. Thermogravimetric analysis (TGA) of aldehyde – amine aqueous reaction

326

products under an inert atmosphere (31) has shown that peak thermal breakdown depends on the

327

reaction system, but occurs between 180 and 370 °C. Thermograms are shown in Figure S2,

328

normalized at the starting and ending experimental temperatures (105 and 450 °C, respectively).

329

Thermograms of fine continental organic aerosol have two thermal breakdown peaks, the first

330

attributed to volatiles and the second (at ~260 °C) to oligomerized material (64). A study of

331

rural / coastal WSOC also observed a third region of thermal breakdown (near 500 °C) attributed

332

to stable aromatics (65). Volatilization of aldehyde-amine reaction mixtures occurs over a range

333

roughly consistent with organic aerosol oligomers, but not with stable aromatic WSOC species.

334 335

Surface Activity. HULIS have been found to be surface active (4), especially the polyacidic

336

fraction (66). The uptake by AS aerosol of MG or acetaldehyde from the gas phase at 65% RH

337

enhances aerosol CCN activity (67), an effect attributed to surface activity of the aldehydes

ACS Paragon Plus Environment

15

Environmental Science & Technology

Page 16 of 31

338

themselves. However, while GX and its reaction products are not surface active (23), the

339

combination of reactions involving GX and MG would still be expected to produce some

340

surface-active products.

341 342

The notable similarity between atmospheric HULIS BrC extracts and the products of aqueous-

343

phase aldehyde–amine indicates that these reactions are plausibly significant sources of

344

atmospheric HULIS BrC. Aerosol-based kinetic studies under atmospheric conditions will be

345

necessary to quantitate the size of this source, however. The similarity across a wide range of

346

optical, physical, and chemical parameters summarized above suggests that the practice of using

347

aldehyde + amine reactions to produce “lab-generated” nitrogen-containing HULIS BrC material

348

(when extracts of atmospheric aerosol are not available in adequate quantities) is reasonable.

349 350

Acknowledgments and Funding Sources

351

This work was funded by NSF grants AGS-1129002 and AGS-1523178. LNH thanks Lynn M.

352

Russell for assistance with Beamline 5.3.2 and David Kilcoyne, beamline scientist, for access to

353

the Advanced Light Source. LNH was funded in part by Research Corporation grant 22473 and

354

the Barbara Stokes Dewey foundation.

355 356 357 358

ACS Paragon Plus Environment

16

Page 17 of 31

Environmental Science & Technology

359

Supporting Information

360

Additional details regarding STXM-NEXAFS analysis are provided in the SI. We also provide

361

additional discussion of the shape of UV/visible absorption spectra, specific functional group

362

identification of FTIR spectra, and additional discussion of hygroscopicity measurements. Figure

363

S1 contains the hygroscopicity parameter κ for samples in this study and in other atmospheric

364

HULIS measurements. Figure S2 shows the results of thermogravimetric analysis (TGA) for

365

amine-aldehyde products compared with rural PM1 aerosol. Figure S3 shows the N/C and O/C

366

elemental ratios for atmospheric HULIS and methylglyoxal-amine reaction product material.

367

This information is available free of charge via the Internet at http://pubs.acs.org.

368 369 370 371

ACS Paragon Plus Environment

17

Environmental Science & Technology

Page 18 of 31

500 450 300

350

250

300 200 250

HCHO + AS

Glyox + AS 150

500 450

100

400

Fluorescence Intensity

Emission Wavelength (nm)

400

50

350 300

0

250

Glyox + Glycine 250

300 350 400 250 Excitation Wavelength (nm)

Urban WSOC (Duarte et al., 2004) 300

350

400

372 373

Figure 1: Fluorescence maps of 0.25 M room-temperature pH 4 reaction mixtures.

374

Formaldehyde-AS after 6 d, (top left); glyoxal-AS after 4 h (top right); glyoxal-glycine after 3 d

375

(bottom left, 1 nm exc. bandwidth); and WSOC extract of urban/coastal particles collected in

376

Portugal (bottom right, ref (37) reprinted with kind permission from Springer Science+Business

377

Media).

378

wavelengths) between t = 2 and 6 d and grows stronger to reach maximum intensity at t = 14 d.

The fluorescence signal of formaldehyde-AS mixtures appears (at the same peak

379

ACS Paragon Plus Environment

18

Page 19 of 31

Environmental Science & Technology

380 381

Figure 2: UV-visible absorbance spectra of WSOC collected in Decatur, Georgia (ref (5)) on

382

biomass-burning-influenced (thin green) and non-BB-influenced days (thick green line),

383

compared with 0.25 M aqueous aldehyde-amine mixtures after 3 – 7 days of reaction at 22 C and

384

pH 4 (reference (17)).

385

glycolaldehyde reactions. Bold dotted lines are linear fits to geometric averages of absorbance in

386

all reactions involving a given aldehyde. For glyoxal, only reaction mixtures including amines

387

were averaged.

Red:

glyoxal reactions;

black: methylglyoxal reactions;

blue:

ACS Paragon Plus Environment

19

Environmental Science & Technology

Page 20 of 31

388 389

Figure 3. Complex indices of refraction measured between 532 and 550 nm for aldehyde-amine

390

reaction products (colored diamonds, ref (45)), atmospheric HULIS samples, and fulvic acid

391

standards (labeled). Vertical dashed lines: absorption components, ref (68). Organic acids, ref

392

(69); smoke influence (70); HULIS extracted from Amazonian biomass burning aerosol, ref (71);

393

Suwanee River fulvic acid, HULIS pollution and HULIS smoke, ref (72); BrC spheres, ref (73);

394

and soot, ref (74).

395 396

ACS Paragon Plus Environment

20

Page 21 of 31

Environmental Science & Technology

397 398

Figure 4: FTIR-ATR spectra of dried aldehyde – amino acid droplets and atmospheric HULIS.

399

Before drying, 1 µL droplets contained 0.05 M glyoxal with 0.1 M glycine, serine, or arginine-

400

HCl, or 0.05 M methylglyoxal with 0.1 M arginine-HCl. Dried HULIS extracted from urban

401

aerosol (Guangzhou, China) using HRB sorbent and 2% v/v NH3 / MeOH eluent, pressed into a

402

KBr pellet (black line, ref (46) reprinted with permission from Elsevier); and HULIS extracted

403

from urban aerosol (Copenhagen, Denmark) dried on a CaF2 window (blue line, ref (39)

404

reprinted with kind permission from Springer Science+Business Media) are also shown.

ACS Paragon Plus Environment

21

Environmental Science & Technology

Page 22 of 31

405 1

406

Figure 5:

H NMR spectra of (a) methylglyoxal – methylamine reaction mixture, (b) HULIS

407

extracted from NIST 1648 urban dust sample (ref (3) reprinted with kind permission from

408

Springer Science+Business Media), and PMF factor spectra from PM1 aerosol collected in

409

Cabauw, Netherlands (ref (47)). Top two spectra include HDO solvent peak near 4.7 ppm.

410

ACS Paragon Plus Environment

22

Page 23 of 31

Environmental Science & Technology

411 412

Figure 6: Overlay of 13C NMR spectra produced in individual reactions between GX or MG and

413

amino acids (glycine, serine, arginine, or ornithine) (narrow black and gray peaks), along with.

414

CPMAS 13C solid-state NMR spectra of WSOC extracted from autumn (red) and summer (black)

415

rural aerosol (Portugal, ref (48) reprinted with permission from Elsevier). Inset: GX – arginine

416

reaction product.

ACS Paragon Plus Environment

23

Environmental Science & Technology

Page 24 of 31

417 418

Figure 7: Positive ion mode electrospray mass spectra (ESI-MS) for dried and redissolved

419

glycolaldehyde – methylamine mixture (red), compared with negative ion mode electrospray

420

mass spectra for HULIS extracts (from ref (39)) from Copenhagen (“urban,” black), Melpitz,

421

Germany (“rural,” green), and Storm Peak Laboratory, Colorado (“remote,” blue), normalized by

422

air sampling volumes. All ion counts are given as 106.

ACS Paragon Plus Environment

24

Environmental Science & Technology

π

π

σ

π

Page 25 of 31

423 424

Figure 8: STXM-NEXAFS carbon K-edge spectra from four GX-methylamine particles are

425

averaged (black trace) and compared with previously published spectra including atmospheric

426

HULIS (teal), fractions F2 (solid grey) and F5 (dashed grey) of Suwanee River fulvic acid (57),

427

and aerosol labeled as “tar balls” collected during biomass burning episodes in Yosemite, CA

428

(pink) and Flagstaff, AZ (green) (36). Vertical lines indicate the approximate location of various

429

functional groups. Spectra from (57) and (36) were reproduced with kind permission from

430

Elsevier and Csiro publishing, respectively.

431 432

ACS Paragon Plus Environment

25

Environmental Science & Technology

433

Page 26 of 31

References

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

1. Denkenberger, K. A.; Moffet, R. C.; Holecek, J. C.; Rebotier, T. P.; Prather, K. A., Realtime, single-particle measurements of oligomers in aged ambient aerosol particles. Environ. Sci. Technol. 2007, 41, (15), 5439-5446. 2. Wang, X.; Gao, S.; Yang, X.; Chen, H.; Chen, J.; Zhuang, G.; Surratt, J. D.; Chan, M. N.; Seinfeld, J. H., Evidence for high molecular weight nitrogen-containing organic salts in urban aerosols. Environ. Sci. Technol. 2010, 44, (12), 4441-4446. 3. Havers, N.; Burba, P.; Lambert, J.; Klockow, D., Spectroscopic characterization of humic-like substances in airborne particulate matter. J. Atmos. Chem. 1998, 29, 45-54. 4. Graber, E. R.; Rudich, Y., Atmospheric HULIS: How humic-like are they? A comprehensive and critical review. Atmos. Chem. Phys. 2006, 6, 729-753. 5. Hecobian, A.; Zhang, X.; Zheng, M.; Frank, N.; Edgerton, E. S.; Weber, R. J., Watersoluble organic aerosol material and the light-absorption characteristics of aqueous extracts measured over the southeastern United States. Atmos. Chem. Phys. 2010, 10, 5965-5977. 6. Baduel, C.; Voisin, D.; Jaffrezo, J.-L., Seasonal variations of concentrations and optical properties of water soluble HULIS collected in urban environments. Atmos. Chem. Phys. 2010, 10, 4085-4095. 7. Ramanathan, V.; Li, F.; Ramana, M. V.; Praveen, P. S.; Kim, D.; Corrigan, C. E.; Nguyen, H.; Stone, E. A.; Schauer, J. J.; Carmichael, G. R.; Adhikary, B.; Yoon, S. C., Atmospheric brown clouds: hemispherical and regional variations in long-range transport, absorption, and radiative forcing. J. Geophys. Res. 2007, 112, (D12), D22S21/1-D22S21/26. 8. Bahadur, R.; Praveen, P. S.; Xu, Y.; Ramanathan, V., Solar absorption by elemental and brown carbon determine from spectral observations. Proc. Natl. Acad. Sci. (USA) 2012, 109, (43), 17366-17371. 9. Feng, Y.; Ramanathan, V.; Kotamarthi, V. R., Brown carbon: a significant atmospheric absorber of solar radiation? Atmos. Chem. Phys. 2013, 13, (17), 8607-8621. 10. Schill, G. P.; De Haan, D. O.; Tolbert, M. A., Heterogeneous ice nucleation on simulated secondary organic aerosol. Environ. Sci. Technol. 2014, 48, 1675-1682. 11. Koop, T.; Bookhold, J.; Shiraiwa, M.; Pöschl, U., Glass transition and phase state of organic compounds: Dependency on molecular properties and implications for secondary organic aerosols in the atmosphere. Phys. Chem. Chem. Phys. 2011, 13, 19238-19255. 12. Laskin, A.; Laskin, J.; Nizkorodov, S. A., Chemistry of Atmospheric Brown Carbon. Chem. Rev. 2015, 115, 4335-4382. 13. Bones, D. L.; Henricksen, D. K.; Mang, S. A.; Gonsior, M.; Bateman, A. P.; Nguyen, T. B.; Cooper, W. J.; Nizkorodov, S. A., Appearance of strong absorbers and fluorophores in limonene-O3 secondary organic aerosol due to NH4+-mediated chemical aging over long time scales. J. Geophys. Res. - Atmos. 2010, 115, (D5), D05203/1-14. 14. Noziere, B.; Dziedzic, P.; Cordova, A., Formation of secondary light-absorbing "fulviclike" oligomers: a common process in aqueous and ionic atmospheric particles? Geophys. Res. Lett. 2007, 34, L21812.

ACS Paragon Plus Environment

26

Page 27 of 31

474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518

Environmental Science & Technology

15. Noziere, B.; Cordova, A., A kinetic and mechanistic study of the amino acid catalyzed aldol condensation of acetaldehyde in aqueous and salt solutions. J. Phys. Chem. 2008, 112, (13), 2827-2837. 16. Noziere, B.; Dziedzic, P.; Cordova, A., Inorganic ammonium salts and carbonate salts are efficient catalysts for aldol condensation in atmospheric aerosols. Phys. Chem. Chem. Phys. 2010, 12, (15), 3864-3872. 17. Powelson, M. H.; Espelien, B. M.; Hawkins, L. N.; Galloway, M. M.; De Haan, D. O., Brown carbon formation by aqueous-phase aldehyde reactions with amines and ammonium sulfate. Environ. Sci. Technol. 2014, 48, (2), 985-993. 18. Chang, J. L.; Thompson, J. E., Characterization of colored products formed during irradiation of solutions containing H2O2 and phenolic compounds. Atmos. Environ. 2010, 44, 541-551. 19. Tan, Y.; Lim, Y. B.; Altieri, K. E.; Seitzinger, S. P.; Turpin, B. J., Mechanisms leading to oligomers and SOA through aqueous photooxidation: insights from OH radical oxidation of acetic acid and methylglyoxal. Atmos. Chem. Phys. 2012, 12, 801-813. 20. Lim, Y. B.; Tan, Y.; Perri, M. J.; Seitzinger, S. P.; Turpin, B. J., Aqueous chemistry and its role in secondary organic aerosol (SOA) formation. Atmos. Chem. Phys. 2010, 10, 1052110539. 21. Ortiz-Montalvo, D. L.; Schwier, A. N.; LIm, Y. B.; McNeill, V. F.; Turpin, B. J., Volatility of methylglyoxal cloud SOA formed through OH radical oxidation and droplet evaporation. Atmos. Environ. 2016, 130, 145-152. 22. Galloway, M. M.; Chhabra, P. S.; Chan, A. W. H.; Surratt, J. D.; Flagan, R. C.; Seinfeld, J. H.; Keutsch, F. N., Glyoxal uptake on ammonium sulphate seed aerosol: reaction products and reversibility of uptake under dark and irradiated conditions. Atmos. Chem. Phys. 2009, 9, 3331-3345. 23. Shapiro, E. L.; Szprengiel, J.; Sareen, N.; Jen, C. N.; Giordano, M. R.; McNeill, V. F., Light-absorbing secondary organic material formed by glyoxal in aqueous aerosol mimics. Atmos. Chem. Phys. 2009, 9, 2289-2300. 24. Sareen, N.; Schwier, A. N.; Shapiro, E. L.; Mitroo, D.; McNeill, V. F., Secondary organic material formed by methylglyoxal in aqueous aerosol mimics. Atmos. Chem. Phys. 2010, 10, 997-1016. 25. Yu, G.; Bayer, A. R.; Galloway, M. M.; Korshavn, K. J.; Fry, C. G.; Keutsch, F. N., Glyoxal in aqueous ammonium sulfate solutions: products, kinetics, and hydration effects. Environ. Sci. Technol. 2011, 45, 6336-6342. 26. Kampf, C. J.; Jakob, R.; Hoffmann, T., Identification and characterization of aging products in the glyoxal/ammonium sulfate system -- implications for light-absorbing material in atmospheric aerosols. Atmos. Chem. Phys. 2012, 12, 6323-6333. 27. Sareen, N.; Moussa, S. G.; McNeill, V. F., Photochemical Aging of Light-Absorbing Secondary Organic Aerosol Material. The Journal of Physical Chemistry A 2013, 117, (14), 2987-2996. 28. De Haan, D. O.; Tolbert, M. A.; Jimenez, J. L., Atmospheric condensed-phase reactions of glyoxal with methylamine. Geophys. Res. Lett. 2009, 36, L11819. 29. De Haan, D. O.; Hawkins, L. N.; Kononenko, J. A.; Turley, J. J.; Corrigan, A. L.; Tolbert, M. A.; Jimenez, J. L., Formation of nitrogen-containing oligomers by methylglyoxal and amines in simulated evaporating cloud droplets. Environ. Sci. Technol. 2011, 45, (3), 984-991.

ACS Paragon Plus Environment

27

Environmental Science & Technology

519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564

Page 28 of 31

30. Woo, J. L.; Kim, D. D.; Schwier, A. N.; Li, R.; McNeill, V. F., Aqueous aerosol SOA formation: impact on aerosol physical properties. Faraday Discuss. 2013. 31. Galloway, M. M.; Powelson, M. H.; Sedehi, N.; Wood, S. E.; Millage, K. D.; Kononenko, J. A.; Rynaski, A. D.; De Haan, D. O., Secondary organic aerosol formation during evaporation of droplets containing atmospheric aldehydes, amines, and ammonium sulfate. Environ. Sci. Technol. 2014, 48, 14417-14425. 32. Sedehi, N.; Takano, H.; Blasic, V. A.; Sullivan, K. A.; De Haan, D. O., Temperature- and pH-dependent aqueous-phase kinetics of the reactions of glyoxal and methylglyoxal with atmospheric amines and ammonium sulfate. Atmos. Environ. 2013, 77, 656-663. 33. Takahama, S.; Gilardoni, S.; Russell, L. M.; Kilcoyne, A. L. D., Classification of multiple types of organic carbon composition in atmospheric particles by scanning transmission X-ray microscopy analysis. Atmos. Environ. 2007, 41, 9435-9451. 34. Takahama, S.; Liu, S.; Russell, L. M., Coatings and clusters of carboxylic acids in carbon-containing atmospheric particles from spectromicroscopy and their implications for cloud-nucleating and optical properties. J. Geophys. Res. 2010, 115, D01202. 35. Hawkins, L. N.; Russell, L. M., Polysaccharides, proteins, and phytoplankton fragments: Four chemically distinct types of marine primary organic aerosol classified by single particle spectromicroscopy. Adv Meteorol. 2010, 2010, 612132, 14 pp. 36. Hopkins, R. J.; Tivanski, A. V.; Marten, B. D.; Gilles, M. K., Chemical bonding and structural information of black carbon reference materials and individual carbonaceous atmospheric aerosols. J. Aerosol Sci. 2007, 38, 573-591. 37. Duarte, R. M. B. O.; Pio, C. A.; Duarte, A. C., Synchronous scan and excitation-emission matrix fluorescence spectroscopy of water-soluble organic compounds in atmospheric aerosols. J. Atmos. Chem. 2004, 48, (2), 157-171. 38. Nguyen, T. B.; Lee, P. B.; Updyke, K. M.; Bones, D. L.; Laskin, J.; Laskin, A.; Nizkorodov, S. A., Formation of nitrogen- and sulfur-containing light-absorbing compounds accelerated by evaporation of water from secondary organic aerosols. J. Geophys. Res. 2012, 117, D01207. 39. Kristensen, T. B.; Du, L.; Nguyen, Q. T.; Nøjgaard, J. K.; Bender Koch, C.; Faurskov Nielsen, O.; Hallar, A. G.; Lowenthal, D. H.; Nekat, B.; van Pinxteren, D.; Herrmann, H.; Glasius, M.; Kjaergaard, H. G.; Bilde, M., Chemical properties of HULIS from three different environments. J. Atmos. Chem. 2015, 72, 65-80. 40. Phillips, S. M.; Smith, G. D., Further evidence for charge transfer complexes in brown carbon aerosols from excitation-emission matrix fluorescence spectroscopy. J. Phys. Chem. 2015, 119, (19), 4545-4551. 41. Phillips, S. M.; Smith, G. D., Light absorption by charge transfer complexes in brown carbon aerosols. Environ. Sci. Technol. Lett. 2014, 1, 382-386. 42. Rincón, A. G.; Guzmán, M. I.; Hoffmann, M. R.; Colussi, A. J., Thermochromism of model organic aerosol matter. J. Phys. Chem. Lett. 2010, 1, 368-373. 43. Srinivas, B.; Rastogi, N.; Sarin, M. M.; Singh, A.; Singh, D., Mass absorption efficiency of light absorbing organic aerosols from source region of paddy-residue burning emissions in the Indo-Gangetic Plain. Atmos. Environ. 2015, 125, (B), 360-370. 44. Massabò, D.; Caponi, L.; Bernardoni, V.; Bove, M. C.; Brotto, P.; Calzolai, G.; Cassola, F.; Chiari, M.; Fedi, M. E.; Fermo, P.; Giannoni, M.; Lucarelli, F.; Nava, S.; Piazzalunga, A.; Valli, G.; Vecchi, R.; Prati, P., Multi-wavelength optical determination of black and brown carbon in atmospheric aerosols. Atmos. Environ. 2015, 108, 1-12.

ACS Paragon Plus Environment

28

Page 29 of 31

565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610

Environmental Science & Technology

45. Zarzana, K. J.; De Haan, D. O.; Freedman, M. A.; Hasenkopf, C. A.; Tolbert, M. A., Optical Properties of the Products of α-Dicarbonyl and Amine Reactions in Simulated Cloud Droplets. Environ. Sci. Technol. 2012, 46, (9), 4845-4851. 46. Fan, X.; Song, J.; Peng, P., Comparative study for separation of atmospheric humic-like substance (HULIS) by ENVI-18, HLB, XAD-8 and DEAE sorbents: Elemental composition, FT-IR, 1H NMR and off-line thermochemolysis with tetramethylammonium hydroxide (TMAH). Chemosphere 2013, 93, 1710-1719. 47. Paglione, M.; Kiendler-Scharr, A.; Mensah, A. A.; Finessi, E.; Giulianelli, L.; Sandrini, S.; Facchini, M. C.; Fuzzi, S.; Schlag, P.; Piazzalunga, A.; Tagliavini, E.; Henzing, J. S.; Decesari, S., Identification of humic-like substances (HULIS) in oxygenated organic aerosols using NMR and AMS factor analyses and liquid chromatographic techniques. Atmos. Chem. Phys. 2014, 14, (1), 25-45. 48. Duarte, R. M. B. O.; Pio, C. A.; Duarte, A. C., Spectroscopic study of the water-soluble organic matter isolated from atmospheric aerosols collected under different atmospheric conditions. Anal. Chim. Acta 2005, 530, 7-14. 49. De Haan, D. O.; Corrigan, A. L.; Smith, K. W.; Stroik, D. R.; Turley, J. T.; Lee, F. E.; Tolbert, M. A.; Jimenez, J. L.; Cordova, K. E.; Ferrell, G. R., Secondary organic aerosol-forming reactions of glyoxal with amino acids. Environ. Sci. Technol. 2009, 43, (8), 2818-2824. 50. Salma, I.; Ocskay, R.; Chi, X.; Maenhaut, W., Sampling artefacts, concentration and chemical composition of fine water-soluble organic carbon and humic-like substances in a continental urban atmospheric environment. Atmos. Environ. 2007, 41, 4106-4118. 51. Krivácsy, Z.; Gelencsér, A.; Kiss, G.; Mészáros, E.; Molnár, A.; Hoffer, A.; Mészáros, T.; Sárvári, Z.; Temesi, D.; Varga, B.; Baltensperger, U.; Nyeki, S.; Weingartner, E., Study on the chemical character of water soluble organic compounds in fine atmospheric aerosol at the Jungfraujoch. J. Atmos. Chem. 2001, 39, 235-259. 52. Kiss, G.; Varga, B.; Galambos, I.; Ganszky, I., Characterization of water-soluble organic matter isolated from atmospheric fine aerosol. J. Geophys. Res. 2002, 107, (D21), ICC 1-1 - ICC 1-8. 53. Limbeck, A.; Kulmala, M.; Puxbaum, H., Secondary organic aerosol formation in the atmosphere via heterogeneous reaction of gaseous isoprene on acidic particles. Geophys. Res. Lett. 2003, 30, (19). 54. Kiss, G.; Tombácz, E.; Varga, B.; Alsberg, T.; Persson, L., Estimation of the average molecular weight of humic-like substances isolated from fine atmospheric aerosol. Atmos. Environ. 2003, 37, 3783-3794. 55. Nguyen, Q. T.; Kristensen, T. B.; Hansen, A. M. K.; Skov, H.; Bossi, R.; Massling, A.; Sørensen, L. L.; Bilde, M.; Glasius, M.; Nøjgaard, J. K., Characterization of humic-like substances in Arctic aerosols. J. Geophys. Res. Atmos. 2014, 119, (8), 5011-5027. 56. Laskin, A.; Smith, J. S.; Laskin, J., Molecular characterization of nitrogen-containing organic compounds in biomass burning aerosols using high-resolution mass spectrometry. Environ. Sci. Technol. 2009, 43, (10), 3764-3771. 57. Zelenay, V.; Huthwelker, T.; Křepelová, A.; Rudich, Y.; Ammann, M., Humidity driven nanoscale chemical separation in complex organic matter. Environ. Chem. 2011, 8, (4), 450-460. 58. Myneni, S. C. B., Soft X-ray spectroscopy and spectromicroscopy studies of organic molecules in the environment. Rev Mineral Geochem 2002, 49, (1), 485-579. 59. Fors, E. O.; Rissler, J.; Massling, A.; Svenningsson, B.; Andreae, M. O.; Dusek, U.; Frank, G. P.; Hoffer, A.; Bilde, M.; Kiss, G.; Janitsek, S.; Henning, S.; Facchini, M. C.;

ACS Paragon Plus Environment

29

Environmental Science & Technology

611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653

Page 30 of 31

Decesari, S.; Swietlicki, E., Hygroscopic properties of Amazonian biomass burning and European background HULIS and investigation of their effects on surface tension with two models linking H-TDMA to CCNC data. Atmos. Chem. Phys. 2010, 10, 5625-5639. 60. Dinar, E.; Taraniuk, I.; Graber, E. R.; Anttila, T.; Mentel, T. F.; Rudich, Y., Hygroscopic growth of atmospheric and model humic-like substances. J. Geophys. Res. 2007, 112, D05211. 61. Hawkins, L. N.; Baril, M. J.; Sedehi, N.; Galloway, M. M.; De Haan, D. O.; Schill, G. P.; Tolbert, M. A., Formation of semi-solid, oligomerized aqueous SOA: Lab simulations of cloud processing. Environ. Sci. Technol. 2014, 48, (4), 2273-2280. 62. Petters, M. D.; Kreidenweis, S., A single parameter representation of hygroscopic growth and cloud condensation nucleus activity. Atmos. Chem. Phys. 2007, 7, 1961-1971. 63. Prenni, A. J.; Petters, M. D.; Kreidenweis, S. M.; DeMott, P. J.; Ziemann, P. J., Cloud droplet activation of secondary organic aerosol. J. Geophys. Res. 2007, 112, D10223. 64. Gelencsér, A.; Hoffer, A.; Molnár, A.; Krivácsy, Z.; Kiss, G.; Mészáros, E., Thermal behaviour of carbonaceous aerosol from a continental background site. Atmos. Environ. 2000, 34, 823-831. 65. Duarte, R. M. B. O.; Duarte, A. C., Thermogravimetric characteristics of water-soluble organic matter from atmospheric aerosols collected in a rural–coastal area. Atmos. Environ. 2008, 42, (27), 6670-6678. 66. Facchini, M. C.; Decesari, S.; Mircea, M.; Fuzzi, S.; Loglio, G., Surface tension of atmospheric wet aerosol and cloud/fog droplets in relation to their organic carbon content and chemical composition. Atmos. Environ. 2000, 34, 4853-4857. 67. Sareen, N.; Schwier, A. N.; Lathem, T. L.; Nenes, A.; McNeill, V. F., Surfactants from the gas phase may promote cloud droplet formation. Proc. Natl. Acad. Sci. (USA) 2013, 110, (8), 2723-2728. 68. Kirchstetter, T. W.; Novakov, T.; Hobbs, P. V., Evidence that the spectral dependence of light absorptionby aerosols is affected by organic carbon. J. Geophys. Res. - Atmos. 2004, 109, D21208. 69. Lund Myhre, C. E.; Nielsen, C. J., Optical properties in the UV and visible spectral region of organic acids relevant to tropospheric aerosols. Atmos. Chem. Phys. 2004, 4, 17591769. 70. Guyon, P.; Boucher, O.; Graham, B.; Beck, J.; Mayol-Bracero, O. L.; Roberts, G. C.; Maenhaut, W.; Artaxo, P.; Andreae, M. O., Refractive index of aerosol particles over the Amazon tropical forest during LBA-EUSTACH 1999. J. Aerosol Sci. 2003, 34, (7), 883-907. 71. Hoffer, A.; Gelencser, 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, 3563-3570. 72. Dinar, E.; Abo Riziq, 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. 2007, 137, 279-295. 73. Alexander, D. T. L.; Crozier, P. A.; Anderson, J. R., Brown carbon spheres in East Asian outflow and their optical properties. Science 2008, 321, 833-836. 74. Bond, T. C.; Bergstrom, R. W., Light absorption by carbonaceous particles: an investigative review. Aerosol Sci. Technol. 2006, 40, 27-67.

654 655

ACS Paragon Plus Environment

30

istry em31 Page Environmental 31 hof Science & Technology lard c with: cts Mail produ /visible V forms less U ature

✓ Faebsorptiont fluorescence ican ✓ SHigygnrifoscopaictmgorsopwhtheric ar to ✓ sHimUilLIS ers ✓ Oligom refractive indices r IS Simila C and HUL SO ✓ to WACS and NMR, wn Plus Environment TIR,Paragon rF to bro ectra

Simila

✓ Xc-rarabyosnpand HULIS