Optical Properties of Airborne Soil Organic Particles - ACS Earth and

Sep 8, 2017 - (17, 48) EELS measurements were performed using a Gatan EELS spectrometer controlled by Digital Micrograph software to collect high-ener...
0 downloads 10 Views 3MB Size
Subscriber access provided by UNIV OF WESTERN ONTARIO

Article

Optical Properties of Airborne Soil Organic Particles Daniel P. Veghte, Swarup China, Johannes Weis, Libor Kovarik, Mary K. Gilles, and Alexander Laskin ACS Earth Space Chem., Just Accepted Manuscript • DOI: 10.1021/ acsearthspacechem.7b00071 • Publication Date (Web): 08 Sep 2017 Downloaded from http://pubs.acs.org on September 11, 2017

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

ACS Earth and Space Chemistry 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 38

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

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

ACS Earth and Space Chemistry

Optical Properties of Airborne Soil Organic Particles Daniel P. Veghte1, Swarup China1, Johannes Weis2,3, Libor Kovarik1, Mary K. Gilles2, Alexander Laskin4,* 1

William R. Wiley Environmental Molecular Sciences Laboratory, Pacific Northwest National Laboratory, Richland, Washington 99354, USA. 2 Chemical Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA. 3 Department of Chemistry, University of California, Berkeley, California 94720, USA. 4 Department of Chemistry, Purdue University, West Lafayette, IN 47907-2084 USA. Manuscript in Preparation for: ACS Earth and Space Chemistry

*Correspondence: [email protected]

Keywords: Airborne soil organic particles, optical constants, electron energy loss spectroscopy, scanning transmission electron microscopy, refractive index

1 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

Page 2 of 38

47 48 49

Abstract:

50

organic particles (ASOP). The Chemical composition of ASOP include macromolecules such as

51

polysaccharides, tannins, and lignin (derived from degradation of plants and biological

52

organisms), which determine light absorbing (brown carbon) particle properties. Optical

53

properties of ASOP were inferred from the quantitative analysis of the electron energy-loss

54

spectra acquired over individual particles using transmission electron microscopy. The optical

55

constants of ASOP are compared with those measured for laboratory generated particles

56

composed of Suwanee River Fulvic Acid (SRFA) reference material, which is used as a

57

laboratory surrogate of ASOP. The chemical composition of the particles was analyzed using

58

energy dispersive x-ray spectroscopy, electron energy-loss spectroscopy, and synchrotron-based

59

scanning transmission x-ray microscopy with near edge x-ray absorption fine structure

60

spectroscopy. ASOP and SRFA exhibit similar carbon composition, with minor differences in

61

other elements present. When ASOP are heated to 350 °C their absorption increases as a result of

62

pyrolysis and partial volatilization of semi-volatile organic constituents. The retrieved refractive

63

index (RI) at 532 nm of SRFA particles, ASOP, and heated ASOP were 1.22-0.07i, 1.29-0.07i,

64

and 1.90-0.38i, respectively. Retrieved imaginary part of the refractive index of SRFA particles

65

derived from EELS measurements was higher and the real part was lower compared to data from

66

more common optical methods. Therefore, corrections to the EELS data are needed for

67

incorporation into models. These measurements of ASOP optical constants confirm that they

68

have properties characteristic of atmospheric brown carbon and therefore their potential effects

69

on the radiative forcing of climate need to be assessed in atmospheric models.

The impact of water droplets on soils has recently been found to drive emissions of airborne soil

70

2 ACS Paragon Plus Environment

Page 3 of 38

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

71

72 73

Introduction: Atmospheric aerosols affect the Earth’s climate directly by scattering and absorbing solar

74

and terrestrial radiation. Aerosol optical properties depend on the chemical composition, sizes,

75

shapes, mixing states, and refractive indices of individual particles as well as aerosol

76

concentrations.1-6 An additional indirect effect of aerosols occurs when they modify optical and

77

microphysical properties of clouds where particles act as ice and cloud condensation nuclei.7

78

Light absorbing particles in the troposphere are of particular interest due to their warming

79

effect.6 The two major classes of carbonaceous absorbing particles are black carbon (BC) and

80

brown carbon (BrC). While BC dominates the absorption, BrC contributes to between 18-56% of

81

the carbonaceous absorption of solar radiation globally depending on the wavelength.8-10 BrC

82

particles have a broad range of chemical compositions and morphologies ranging from low

83

viscosity secondary organic material to glassy tar balls and airborne soil organic particles.11-14

84

BrC is composed of various types of organic components that absorb light at variable

85

wavelengths but are generally strong absorbers in the UV and short wavelength of the visible

86

range.15-16 While BrC light absorption is enhanced at shorter wavelengths, BC absorbs strongly

87

over all wavelengths.16-17 For example, tar balls are absorbing BrC species that exhibit a large

88

variability in their optical properties with refractive indices (RI) with a range of real part from

89

1.56-1.88 and imaginary part from 0.002i to 0.27i. 13-14 Additionally, both BC and BrC particles

90

can be coated with organic components that increase the absorption due to a lensing effect.18-21

91

Field measurements use the absorption Ångström exponent (AAE) to describe the wavelength

92

dependence of the light-absorption by aerosols with respect to the empirical power law of ~λ-AAE,

3 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

Page 4 of 38

93

where λ is wavelength. The broad range of AAE values reported in the literature (1.6-11)15, 22 is

94

an inherent result of the large variability in the BrC chemical composition resulting in substantial

95

uncertainty in quantitative assessment of optical properties of aerosol containing BrC and BC

96

components.23-24 Most models assume that the majority of absorbance by atmospheric

97

carbonaceous material is due to highly absorbing BC and do not account for BrC present in the

98

atmosphere, which could be considerable.25

99

Airborne soil organic particles (ASOP) were recently recognized as a distinct particle

100

type with a composition similar to soil organic matter (SOM). Their substantial contribution (up

101

to 60%, by number), detected at the measurement site located in the Southern Great Plains

102

(Oklahoma, USA), was attributed to atmosphere–land interactions during rainfall.12 It is

103

suggested that these particles may be prevalent in certain geographic areas where open soils are

104

exposed to precipitation events such as agricultural fields and grasslands.12, 26 SOM is a large

105

source of terrestrial carbon and produced from plant litter being decomposed in biochemical

106

processes.27 The suggested mechanism of ASOP formation is through incorporation of small air

107

bubbles in the aqueous layers of wet soils followed by bursting of the bubbles at the air-water

108

interface releasing a fine mist of SOM-containing microdroplets that upon later dehydration form

109

solid ASOP.28 Since the composition of ASOP and SOM are similar, the optical properties of

110

ASOP could potentially be approximated by SOM proxies such as Suwanee River Fulvic Acid

111

(SRFA) standard material.29-32

112

Understanding the broad-band optical properties of individual atmospheric particles is

113

essential for estimating direct effects of aerosols on climate. Methods used to measure the

114

extensive optical properties of airborne ensembles of particles include: nephelometers, Fourier

115

transform infrared spectrometer, cavity ring-down (CRDS) and photoacoustic spectrometer

4 ACS Paragon Plus Environment

Page 5 of 38

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

116

(PAS).33 Nephelometers and mid-IR spectrometers have a high detection limit and are limited to

117

high concentrations of polydisperse samples.33-34 CRDS and PAS have lower detection limits

118

(approximately 0.4 Mm-1 35 and 0.8 Mm-1 36 respectively), and thus can only measure the optical

119

properties of ensemble of particles at limited wavelengths. 29-30, 34 Additionally, filter based

120

methods, such as particle soot absorption photometers and aethalometers can measure light

121

absorption by bulk particle samples deposited on the filters but can have significant scattering

122

artifacts that require the use of correction factors.34, 37 However, all of these methods require

123

either large ensembles of airborne particles or bulk samples.

124

Extensive optical properties of particles depend on their size, morphology, and

125

composition of each of the particle components. The RI derived using an effective medium

126

approximation of a particle is a function of the chemical composition, phase, and mixing state

127

that can be uniquely assessed using electron microscopy and chemical imaging methods.38-39 In

128

certain cases, they can probe optical properties of individual particles of interest, identifying

129

them from a complex mixture of particles collected on substrates. For example, quantitative

130

analysis of the particle-specific low-loss energy spectra from electron energy-loss spectroscopy

131

(EELS) coupled to the scanning transmission electron microscope (STEM) was previously used

132

to quantify intensive optical properties of individual tar balls from biomass burning on a single

133

particle basis.17, 40 By using the low-loss region of EELS, RI can be retrieved in the 200-1200 nm

134

wavelength range for individual particles, provided that particles were stable under the electron

135

beam and their morphology would allow estimation of thickness along the direction of the

136

incident electron beam. With this method of examining individual particles, field samples

137

containing a mixture of particle types that satisfy these conditions can be analyzed and optical

138

constants can be retrieved for each particle type present in the mixture. Because of the refractory

5 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

139

behavior of ASOP and their stability under the electron beam,12 EELS provides an optimal

140

method to probe the optical properties of ASOP.

Page 6 of 38

141

In this paper, we present the optical properties of individual ASOP retrieved from low-

142

loss EELS data and compare those to the laboratory generated particles composed of Suwanee

143

River Fulvic Acid (SRFA) reference material. SRFA is a well-studied material with RI values

144

reported in literature.29-32, 41 We analyze the chemical composition of ASOP and SRFA particles

145

using scanning transmission x-ray microscopy near edge x-ray absorption fine structure

146

(STXM/NEXAFS), and TEM based methods of energy dispersive x-ray spectroscopy (EDX),

147

and EELS. Both ASOP and SRFA are primarily carbonaceous, however EDX detects some S

148

and Na in SRFA. NEXAFS shows that ASOP have higher levels of carbonization (more

149

intensive sp2 (C=C) peak) compared to the SRFA particles, while SRFA contains more

150

prominent oxygenated functional groups. Consistent with their compositional characteristics,

151

ASOP and SRFA have similar imaginary parts (absorption) of the RI, while the real part

152

(scattering) of the RI is larger for SRFA. In addition, we show that heating the ASOP particles to

153

350 °C leads to charring of their organic content, leading to substantially larger absorption.

154

155

Experimental Methods:

156

Particle Samples

157

Field samples of ASOP were collected from 4/28/2016 (18:30) to 4/29/2016 (04:30),

158

5/5/2016 (08:00) to 5/5/2016 (21:00), and 5/15/2016 (12:00) to 5/15/2016 (16:45) local time at

159

the Southern Great Plains Site (36° 36’ 18” N, 97° 29’ 6” W) in Lamont, Oklahoma. This site is

160

operated by the Atmospheric Radiation Monitoring (ARM) Program of the U. S. Department of

6 ACS Paragon Plus Environment

Page 7 of 38

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

161

Energy (DOE). Particles used for EELS analysis were mainly from the 4/28/2016 sample.

162

Particles were collected onto the 7th stage (D50 cut-off size of 0.32 µm) of a MOUDI impactor

163

preloaded with two types of microscopy substrates: Si3N4 membrane windows (Silson, Ltd) and

164

400 mesh Copper TEM grids coated with either lacey carbon or Type-B carbon film (Ted Pella,

165

Inc). The meteorological conditions and air parcel backward trajectories calculated using the

166

Hybrid Single-Particle Lagrangian Integrated Trajectory (HYSPLIT) model42 during the

167

sampling periods are given in the supplemental file (Figure S1). The backward trajectories

168

indicate the air mass arrived from the north part of the Great Plains and was associated with

169

relatively calm and humid weather with low wind speeds.

170

Aerosolized particles generated from aqueous solutions of Suwanee River Fulvic Acid

171

(1S101F, International Humic Substances Society) were used as a laboratory standard for the

172

retrieval of optical properties using EELS. SRFA particles were generated by nebulization of a 1

173

wt.% solution of SRFA into a nitrogen gas stream at 4 lpm using a medical nebulizer (8900-7-50,

174

Salter Labs, Arvin, CA). The particle-laden stream was dried using a diffusion dryer (TSI

175

306200, Shoreview, MN). The subsequent dry particles were then collected on the microscopy

176

substrates in the same manner as the field collected ASOP.

177

Selected field samples were heated to 350 °C inside the TEM using a furnace-type

178

heating holder (Gatan, Pleasanton CA) to remove volatile and semi-volatile materials. After

179

heating, STEM/EELS analysis was performed over the thermally modified ASOP. To simulate

180

the conditions of the TEM heating holder, for the STXM/NEXAFS analysis, the corresponding

181

field samples were heated under nitrogen in a tube furnace (MTI GSL 1300X) at 20 °C/min until

182

reaching 350 °C where it was held for 5 minutes before cooling at 20°C/min to room

183

temperature.

7 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

184 185

Page 8 of 38

Imaging and elemental analysis: Particle samples were first imaged by scanning electron microscopy (SEM; Quanta 3D,

186

FEI, Hillsboro, OR) operated at 20 kV. These particles were imaged both orthogonally to the

187

incident electron beam and at a 75° tilt angle to image the contact area between particle and

188

substrate.12 Elemental composition of single particles was measured using a 10 mm2 Si(Li)

189

energy dispersive x-ray (EDX) detector (EDAX Genesis, Mahwah, NJ) interfaced with the SEM

190

and operated at 20 kV.43-44 Samples with a high abundance of ASOP were selected for further

191

analysis by STEM and STXM.

192

An aberration corrected STEM (FEI Titan 80-300 STEM, Hillsboro, OR) operated at 80

193

kV was used to image single particles and the electron detection was performed using the high-

194

angle annular dark-field (HAADF) detector to probe the internal composition of the particles

195

based on the z-contrast. High-loss dual dispersion range EELS (Gatan, Pleasanton, CA) was used

196

to collect maps in STEM mode with 0.25 eV energy resolution in the ranges of: -50 to 500 eV

197

for the zero-loss peak, and 200 to 750 eV to collect data for carbon, oxygen, and nitrogen.

198

Signals were integrated from 280 to 360 eV for the carbon maps and from 525 to 568 eV for the

199

oxygen maps. Elemental maps were collected with a lateral resolution of approximately 3 nm.

200

STXM was used to acquire NEXAFS spectra of individual particles at the carbon k-edge

201

energy range (278-320 eV).12, 45-47 Monochromatic incident photons from the synchrotron source

202

were focused on the sample using a Fresnel zone plate. Sets of raster scan images were acquired

203

over selected fields of view of the particle samples at selected x-ray energies by recording the

204

transmitted light intensity. Spectra of individual particles were then reconstructed based on the

205

Beer Lambert law from the stack images using the signal collected from the areas of individual

206

particles referenced to the background signal from regions without particles. STXM/NEXAFS

8 ACS Paragon Plus Environment

Page 9 of 38

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

207

was performed at the Advanced Light Source synchrotron facility at Lawrence Berkeley National

208

Laboratory on beamlines 11.0.2.2 and 5.3.2.2.12, 45-47

209

EELS acquisition and quantification of optical properties:

210

EELS measurements for optical properties were performed using an aberration corrected

211

STEM operated at 80 kV. The use of an electron beam at 80 kV reduced the Cherenkov radiation

212

effects and also minimized the electron beam induced knock-on damage.17, 48 EELS

213

measurements were performed using a Gatan EELS spectrometer controlled by Digital

214

Micrograph software to collect high-energy resolution, single range dispersion EELS data.

215

Working at 80 kV and using the monochromator decreased the full width at half maximum of the

216

zero-loss peak (ZLP) from 0.8 eV to 0.15-0.2 eV with an energy dispersion of 0.025 eV/channel.

217

Energy-loss spectra were acquired at a collection angle of 45.5 mrad. For each particle, the

218

energy-loss spectrum was obtained by centering the incident electron beam on the middle of the

219

particle allowing the thickness to be estimated as the particle diameter. The electron dose rate to

220

each particle was approximately 0.03 nA, and each spectrum took between 1-2 minutes to

221

collect. After five consecutive EELS spectra were recorded there was no difference observed in

222

the spectra or damage to the particle (Figure S2). Collection of multiple spectra demonstrates

223

that the particles are not beam sensitive as would be observed through changes in subsequent

224

spectra or visible damage to the particles. The reported retrieved RI values were averaged over

225

12 measurements for SRFA particles (RISRFA), 14 measurements for ASOP (RIASOP), and 10

226

measurements for ASOP heated to 350 °C (RIASOPh).

227

The low-loss region (0-10 eV) in the EELS spectrum was used to derive the optical

228

constants of 50-200 nm diameter particles of SRFA, ASOP, and heated ASOP following

229

procedures established by Crozier and co-workers.17, 40 Figure 1 shows an example of the low-

9 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

Page 10 of 38

230

loss EELS spectrum of an individual ASOP, the corresponding fitted ZLP and the single

231

scattering distribution, S(E), used to calculate the particle-specific RI values using the Kramers-

232

Kronig relationship. In the low-loss EELS spectrum, the first peak centered around 0 eV is the

233

ZLP, the second peak around 4-6 eV is the plasmon resonance of the π electrons, while the broad

234

peak around 22 eV is due to the π + σ plasmon peaks.49 Retrieval of the S(E) from experimental

235

EELS data requires removal of the ZLP and correction for the multiple scattering. The S(E)

236

distributions for all analyzed particles are included in the supplemental information (Figure S3).

237

Further discussion for obtaining the S(E) is also in the supplemental information.

238

Monochromatization of the electron beam is essential to produce an electron beam with a narrow

239

energy distribution, which facilitates more accurate removal of the ZLP at lower energies.

240

241

10 ACS Paragon Plus Environment

Page 11 of 38

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

242

Figure 1. Example raw EELS spectra (black) with the extracted single scattering distribution,

243

S(E), (blue) after removal of the zero-loss peak (red) and correction for plural scattering. The

244

shaded energy region corresponds to the retrieval of the RI at 200-1200 nm.

245 246

To retrieve the RI values, the Kramers-Kronig analysis was performed on the extracted

247

S(E) using a modified version of the MATLAB KraKro code (http://tem-eels.com) employing

248

Ritchie’s derivation for thin film approximation.50,51 The thin film approximation has been

249

shown to be effective for particles with diameters of 40-200 nm.17 The low-loss region of the

250

EELS spectrum can be described by the single scattering distribution function S(E) according to

251

equation 151  =



  

252





   1 +   +  

(1)



253

where I0 is the ZLP intensity, v is speed of light, t is the particle diameter, a0 is the Bohr radius,

254

m0 is electron mass, β is the collection angle, θE is the characteristic angle of energy loss, and

255

ε(E)=ε1(E)+iε2(E) is the complex dielectric function for the chemical composition of the probed

256

particle. β is calculated by measuring the diffraction pattern of sapphire in the 110 and 104

257

planes and measuring the diameter of the collection area to obtain the collection angle while θE is

258

calculated based on the energy of the incident electron beam. The SS(E) term represents the

259

energy loss due to surface contributions described by equation 2   =  

260

 

!

"#$ /  



−  ' (  

)

$

'



−  *  +

(2)

261

An iterative method was used to solve eq. (1) for Im[-1/ε(E)]. Initially, the SS(E) term was set to

262

zero and then allowed to vary in subsequent iterations until the solution for Im[-1/ε(E)]

263

converged.50-51 11 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

Page 12 of 38

264

The Kramers-Kronig relationship was then used to retrieve the Re[1/ε(E)] using the causational

265

relationship between Re[1/ε(E)] and Im[-1/ε(E)] where the Cauchy principle part of the integral

266

is denoted by P(eq. 3) 

4

.

,-  = 1 − / 05  

267







 1 23

 1  3  

268

The dielectric function (ε) was then derived from Re[1/ε(E)] and Im[-1/ε(E)] according to

269

equation 4.

270

89:/;'

'= :/;>

(3)

(4)

271

Computation of the dielectric function using the Kramers-Kronig formulation converged within

272

three iterations. Calculations using up to 20 iterations did not significantly deviate from the

273

results with 3 iterations. The real and imaginary parts of the broadband RI(E) = n(E)+ik(E) were

274

then calculated from the dielectric function using equations 5.

275 276

@$ ' ' $ ?@$ '  $ =? A = . . 







(5)

Absorption and scattering cross sections (σabs and σsca) were calculated based on Mie

277

theory and employing retrieved RI. The calculations were performed using BHCOAT code

278

adapted for MATLAB by Mätzler 2002.52 The optical properties of ASOP are presented in terms

279

of the scattering efficiency (Qabs= σabs/πr2), the absorption efficiency (Qsca= σsca/πr2), and the

280

single scattering albedo (SSA= σsca/(σsca+σabs).

281 282 283 284

Results and Discussion: ASOP collected for this study at the SGP site are similar in morphology and composition to those collected previously at SGP.12 For the sample collected on 4/29/2016 and used for 12 ACS Paragon Plus Environment

Page 13 of 38

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

285

further analysis, the HYSPLIT back trajectories indicate that the air parcels spent most of their

286

time over the Great Plains regions where the region is dominated by agriculture and grasslands

287

where ASOP formation is expected (Figure S1).12 For the 5/5/2016 and 5/15/2016 samples, the

288

air mass originated primarily from the North and the previous 72 hours was spent mostly over

289

the Great Plains region. All samples contained an externally mixed particle population and

290

ASOP are observed in all samples. Representative SEM images acquired at 75° tilt angle are

291

shown in Figure 2 with ASOP marked by arrows. Number fractions of ASOP are around 20% for

292

the 4/29/2016 sample, 45% for the 5/5/2016 sample, and 40% for the 5/15/2016 sample. Their

293

sizes are between 100-800 nm with a mean diameter of 500 nm. These particles exhibited a

294

spherical shape with an aspect ratio (width/height) near one, indicative of glassy organic aerosol

295

particles.53 Other collected particles with higher aspect ratios (oblate shapes) were presumably

296

deformed upon impaction on the substrate. These deformed particles with high aspect ratios are

297

often observed in the atmosphere and are usually described as liquid, low viscosity organic

298

products.53 Because ASOP are spherical and resistant to electron beam exposure, they are ideal

299

for the optical constant retrieval based on STEM/EELS measurements.

300

13 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

Page 14 of 38

301 302

Figure 2. Tilted SEM images at 75° of particles collected at the ARM SGP site on a) 4/29/2016,

303

b) 5/5/2016, c) 5/15/2017, and d) laboratory generated SRFA particles. The arrows indicate

304

glassy ASOP. Other particles have relatively low viscosities and presumably flattened during

305

impaction onto the substrate.

306 307

HAADF STEM images in Figure 3 illustrate representative particles of SRFA, ASOP,

308

and heated ASOP. All of these particles are near spherical with an aspect ratio close to one and

309

exhibit no visible signs of damage during electron beam exposure. In addition, as demonstrated

310

by the uniform signal over the particle areas, they do not contain any internal structures. A

311

fraction of ASOP appears to be coated with a thin layer of less viscous (semi-liquid)

312

carbonaceous material adhering to the lacey carbon support film (Fig 3b). After heating to 350

313

°C, ASOP retain their spherical shape with the less viscous material removed (Figure 3c

314

compared to Figure 3b). Heating particles leads to removal of the coating so that the core particle 14 ACS Paragon Plus Environment

Page 15 of 38

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

315

can be analyzed. Literature reports54-55 suggest that heating carbonaceous particles past ~400 °C

316

leads to charring, as is confirmed by the optical data presented below. ASOP and SRFA particles

317

presented here meet the requirements (stability under the electron beam and spherical

318

morphology) for analysis with EELS to retrieve RI values.

319 320

321 322

Figure 3. Representative HAADF STEM images of particles used for RI retrieval: a) SRFA, b)

323

ASOP with a thin coating of semi-liquid carbonaceous material, and c) ASOP heated to 350 °C.

324

Note that the lacey carbon substrate is on the right for all three particles (indicated by the

325

arrows).

326 327

STEM images acquired at the element-specific EELS energies were used to map internal

328

distributions of C and O in ASOP and SRFA particles. Representative images, with a lateral

329

resolution of 3 nm, are shown in Figure 4. The particles also contained nitrogen, but because of

330

its low intensity in the EELS spectra its lateral distribution could not be reliably mapped. STEM

331

imaging of the SRFA particles indicated that the particles are nearly spherical with uniform

332

distributions of C and O. STEM images of ASOP (Figs. 3b and 4d) display visible particle 15 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

Page 16 of 38

333

coatings adhered to the carbon support substrate. Comparison of carbon and oxygen maps (Figs

334

4b, e, h and 4c, f, i, respectively) indicates that particles are mainly carbonaceous. The map of

335

the ASOP oxygen signal (Figure 4f) is enhanced in a thin outer layer (3-6 nm) of the particle.

336

The oxygen-rich coating is indicative of atmospheric processing of the particles through a

337

diffusion limited process.56 Oxygen-rich coatings have been previously observed on tar balls, but

338

much thicker covering approximately the outer 40 nm of the particles.13, 57 In contrast, the maps

339

of heated ASOP show a uniform distribution of carbon and oxygen indicating removal of the

340

oxygen-rich coating.

341

342

16 ACS Paragon Plus Environment

Page 17 of 38

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

343

Figure 4. STEM and EELS (carbon and oxygen) maps of: a-c) SRFA particle, d-f) ASOP, and g-

344

i) heated ASOP. Note that the lacey carbon substrate is on the bottom for the SRFA and on the

345

right for the ASOP and heated ASOP.

346 347

The elemental composition of each particle type was analyzed using SEM/EDS,

348

STEM/EELS, and STXM/NEXAFS (Figure 5). ASOP particles were identified in the tilted SEM

349

images at 75° for further EDS and STXM/NEXAFS experiments. The EDS analysis of each

350

particle type is shown in Figure 5a, where the spectra are scaled to the same height of the oxygen

351

peak to facilitate visual comparison. SRFA is primarily composed of C, N, and O as expected for

352

biodegradation material. SFRA also contains traces of S and Na. ASOP are mainly carbonaceous

353

containing C, N, and O with no additional peaks present from other elements. Similarly, heated

354

ASOP contained only C, N, and O. Figure 5b confirms, by EELS, the presence of C, N, and O in

355

all three samples. The oxygen peak for ASOP was more intense than in the heated ASOP and

356

SRFA particles. ASOP has a relatively higher π* to σ* peak intensity compared to SRFA,

357

consistent with the NEXAFS data shown in Figure 5c. The carbon bonding showed both

358

similarities and differences between particle types. For SRFA particles the most intense peak is

359

RCOOH (288.5 eV) followed by R(C=O)R/C-OH (286.5 eV) and then the C=C sp2 (285.4. eV)

360

peak. The SRFA spectrum shown here is similar to literature reports of other humic acid

361

substances58 which are water-soluble and highly oxidized fractions of the soil organics

362

susceptible to aerosolization by the ‘rain drop mechanism’.28 The NEXAFS spectrum of ASOP is

363

similar to our previously reported spectrum12 and shows a prominent sp2 peak at 285.4 eV and

364

minor peaks of oxygenated functional groups at 286.5 and 288.5 eV. Overall, the spectral

365

features are similar to the ‘free-light’ lower density SOM NEXAFS spectra reported.12 Spectrum

17 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

Page 18 of 38

366

of the heated ASOP shows the sp2 peak is the most prominent peak with smaller oxygen peaks

367

present. This enhancement of the sp2 peak is indicative of charring of the particles which would

368

cause them to be more absorbing and similar to black carbon.54-55, 59 With the heating of the

369

ASOP to remove the coating, it was found that in addition to the coating being removed,

370

composition of the inner particle part has changed as indicated by the enhanced sp2 peak.

371

Relative intensities and shapes of the characteristic features in EDX, EELS, and NEXAFS

372

spectra indicate that ‘free-light’ SOM12 is more similar to ASOP than SRFA.

373

18 ACS Paragon Plus Environment

Page 19 of 38

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

374 375

Figure 5. Elemental comparison of SRFA, ASOP, and ASOP heated to 350 °C particles inferred

376

from: a) EDX spectra normalized to the oxygen counts (Cu and Si are background signals

377

originated from the substrate grid and detector, respectively), b) EELS, and c) STXM/NEXAFS. 19 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

Page 20 of 38

378 379

Figure 6 shows broadband values of RISRFA retrieved over the wavelength range of 200-

380

1200 nm. Previous publications examined the optical constants of atmospheric particles using

381

EELS.17, 40 However, they did not compare measurements on standards to results from other

382

optical methods to determine any systematic errors in the EELS measurements and applications

383

of the Kramers-Kronig analysis. Here, we used SRFA as a standard to compare the single particle

384

EELS method with bulk methods previously employed to retrieve the RISRFA.29-30 The retrieved

385

RISRFA at 390 and 532 nm from this work are compared to those from optical retrieval methods

386

utilizing cavity ring-down spectroscopic methods29-30, 32 in Table 1. Other methods reported the

387

RISRFA for ranges of 315-345 nm,41 360-420 nm,31-32 and single wavelengths of 390 and 532

388

nm.29-30 At 390 nm, the imaginary part of the RISRFA reported here (k=0.08 ) is similar to those

389

reported from other optical methods (k=0.05-0.1).30-32 In contrast, the real part of the RISRFA

390

reported here (1.20) is considerably lower than that measured by optical methods (1.60-1.69).30-32

391

For SRFA we report RISRFA at 532 nm of 1.22-0.07i for the retrieval with EELS, while results

392

from CRDS are between 1.52-0.02i and 1.65-0.02i.29-30 The discrepancy in the CRDS results

393

stems from the use of pulsed CRDS30 compared to continuous wave CRDS29 and a different lot

394

of SRFA which leads to considerable differences in retrieving the optical properties. At these

395

wavelengths, the real part of the reported RISRFA is slightly lower, while the imaginary part of the

396

RISRFA is slightly higher. Overall, the imaginary part of the RISRFA retrieved from EELS is

397

consistent with those from optical methods, while at all wavelengths the real part of the RISRFA is

398

lower. This is similar to previous EELS experiments of polystyrene spheres that were modified

399

using a high electron beam dose and subsequent beam exposure showed no change to the particle

400

morphology .17 Figure S4 shows the comparison of the RI from EELS17 to reported values for

20 ACS Paragon Plus Environment

Page 21 of 38

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

401

polystyrene60 indicating that the real part of the refractive index is substantially lower than the RI

402

retrieved by other methods. Some part of the discrepancy may be due to the modification of the

403

polystyrene spheres with the electron beam. The lower values of the real part of RISRFA retrieved

404

from EELS compared to those derived by the optical methods could also arise from the drier

405

conditions that particles experience under vacuum in the STEM and from using the SRFA

406

standard derived from a purified natural source.29-30 With these caveats in mind, the retrieval of

407

the RI values of atmospheric particles using EELS data, values derived for the RI of the field

408

collected carbonaceous particle presented in this work and in earlier literature17, 40 can be put into

409

a relative perspective.

410

411 412

Figure 6. Comparison of the retrieved broadband refractive index of Suwanee River Fulvic Acid

413

(RISRFA) with data from optical methods.29-32, 41 The real part of the RISRFA(n) is shown in the

21 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

414

upper panel, the imaginary part (k) is shown in the lower panel. Shaded region indicates one

415

standard deviation of the refractive index for all particles measured

Page 22 of 38

416 417

Table 1. Summary of RI values retrieved from STEM/EELS data compared to previous literature

418

values from EELS (*) and optical methods (†) Sample

SRFA

This Study

Literature

(wavelength)

(wavelength)

1.22-0.07i (532 nm)

1.52-0.02i (532 nm) 29, † 1.65-0.02i (532 nm) 30, † 1.602-0.1i (390 nm) 30, †

1.20-0.08i (390 nm)

1.69-0.05i (390 nm) 31, † 1.61-0.05i (390 nm) 32, †

ASOP

1.29-0.07i (532 nm)

-

1.28-0.09i (390 nm) Heated

1.90-0.38i (532 nm)

ASOP

1.81-0.44i (390 nm)

Brown

-

-

1.4-0.1i (550 nm)17,* 1.67-0.27i (550 nm)40,*

Carbon

1.79-0.15i (550 nm)14,† 1.88-0.27i (550 nm)14,† Black Carbon

-

1.8-0.5i (550 nm)17,* 1.95-0.79i (550 nm)40,*

419 420

A comparison of the retrieved RI for SRFA and ASOP is shown in Figure 7. The real part

421

of the RI is larger for ASOP than for SRFA over the entire wavelength region. However, the

422

imaginary parts of the RI for ASOP and SRFA are consistent over the range of 200-1200 nm. For

22 ACS Paragon Plus Environment

Page 23 of 38

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

423

instance, the RIASOP at 532 nm is 1.29-0.07i compared to RISRFA of 1.22-0.07i. Tabulated RIASOP

424

values are included in Table S1. The uncertainty in the retrieved RI SRFA is relatively small with

425

the standard deviation from all particles measured in the real part of 0.02 and in the imaginary

426

part of 0.01 at 532 nm. The uncertainty for the RIASOP values are higher, at 0.06 for the real part

427

and 0.02 for the imaginary part at 532 nm. A larger uncertainty is expected for field-collected

428

samples due to variance in composition. No trend is observed with the variance of the RIASOP

429

with size over the size range of probed particles (50-200 nm). Comparison of the individual

430

single scattering distributions can be found in Figure S3 where the spectra are similar for all

431

particles. The differences between RI SRFA and RIASOP demonstrate that although their absorption

432

properties are comparable, ASOP scatters more efficiently than pure SRFA. Both of these

433

particle types originate from natural organic matter, so similarity in RI would be expected. The

434

major chemical difference between ASOP and SRFA particles is that ASOP has a larger

435

contribution of sp2 to total carbon and an oxygen rich surface.

23 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

Page 24 of 38

436 437

Figure 7. Comparison of the broadband values of RI SRFA (green) and RIASOP (brown). Upper

438

panel shows the real part (n) and lower panel shows the imaginary part (k). Shaded region

439

indicates one standard deviation.

440 441

Figure 8 compares the retrieved values for RIASOP with reported EELS data for field

442

collected tar balls40, amorphous carbon, and graphitic carbon.17 Other BrC particles that have

443

been analyzed previously using EELS to determine the RI were tar balls from biomass burning,

444

collected above the Yellow Sea during the Asian Pacific Regional Aerosol Characterization

445

Experiment.40 Both ASOP and tar balls are derived from plant degradation materials, but with

446

different emission mechanisms. ASOP are emitted from wet soils during rain events.12, 28 Tar

447

balls are generated through the thermal decomposition (pyrolysis) of biomass.13, 61 The RIASOP

24 ACS Paragon Plus Environment

Page 25 of 38

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

448

values (1.29-0.07i at 532 nm) are considerably lower than the RItar balls (1.67-0.27i at 550 nm),40

449

but similar to the amorphous carbon spheres (1.4-0.1i at 532 nm).17 The RItar balls values are

450

slightly lower than the RIASOPh retrieved for heated ASOP and the RIGC of graphitic carbon.17

451

Reported RItar balls values could be slightly higher due to the relativistic effects of using a higher

452

energy electron beam in that study (120 kV versus 80 kV here). Recent laboratory studies show

453

that the retrieved RI values for laboratory generated tar ball proxies are between 1.79-0.15i and

454

1.88-0.27i at 550 nm.14 Additional studies show that tar balls can also have a much lower RI

455

values (1.56-0.02i 13; 1.72-0.008i,1.81-0.006i, 1.75-0.002i).62 When ASOP were heated, the

456

particles charred and had a higher RI than the unheated particles. The retrieved RIASOPh at 532 nm

457

was 1.90-0.38i, which is similar to that reported for graphitic carbon (1.8-0.5i to 1.95-0.79i17 and

458

1.95-0.79i40 at 550 nm). This RIASOPh value is also higher than those reported previously for tar

459

balls. The NEXAFS spectra shown in Figure 5 indicate that when the ASOP were heated, their

460

sp2 peak increased while the oxygenated peaks decreased, indicating charring of particle

461

material. The higher fraction of sp2 carbon present in the heated particles will lead to higher

462

RIASOPh and absorption in the visible light range.63

25 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

Page 26 of 38

463 464

Figure 8. Comparison of the retrieved values of RI SRFA, RIASOP and RIASOPh compared with

465

previous literature employing EELS based method. Upper panel shows the real part (n) and

466

lower panel shows the imaginary part (k) of RI.

467 468

Understanding the optical properties of ASOP is important because they may contribute a

469

significant fraction of the total aerosol present in certain environments.12 Once the RIASOP over a

470

broad wavelength range is retrieved, Mie theory can be used to estimate their contribution to the

471

radiative forcing. Figure 9 shows the scattering efficiency (Qsca), absorption efficiency (Qabs),

472

and single scattering albedo (SSA) calculated for ASOP in the size range of 50-800 nm based on 26 ACS Paragon Plus Environment

Page 27 of 38

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

473

the broadband (200-1200 nm) values of RIASOP. Tabulated data for Qsca, Qabs, and SSA is

474

included in Table S1. Since no substantial difference in the RI was observed for particles

475

between 50-200 nm, the retrieved RI was applied for diameters up to 800 nm to calculate Qsca,

476

Qabs, and the SSA. For particles in the size range of 400-800 nm, the maximum scattering

477

efficiency is approximately at 430 nm. For particle sizes smaller than 400 nm, the scattering

478

efficiency decreases with increasing wavelength. At all particle sizes, the maximum absorption

479

efficiency is at the shortest wavelength (200 nm) and decreases as the wavelength increases. The

480

SSA has a maximum at approximately 600 nm for particles in the size range of 400-800 nm. No

481

maximum absorption was observed for smaller particles that exhibit a decrease in the SSA with

482

increasing wavelength.

483

484 485

Figure 9. Comparison of a) the scattering efficiency (Qsca), b) absorption efficiency (Qabs), and

486

c) single scattering albedo (SSA) of ASOP calculated from the retrieved RIASOP over the

487

wavelength range of 200-1200 nm and particle diameters of 50-800 nm. Note the visible

488

wavelength region is shown by the corresponding colors, while the UV and IR are faded to

489

black.

490 491

27 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

Page 28 of 38

492

The highest prevalence of ASOP was observed on stage 7 of the MOUDI impactor (size

493

range of 0.32-0.56 µm) with most particles having a diameter between 400-600 nm (Figure S5).

494

Stages with either larger or smaller cut-off sizes did not contain ASOP in sufficient quantities to

495

analyze. The size distribution is similar to the previously reported median size of approximately

496

500 nm by Wang et al. 2016 where a Sioutas cascade (SKC Inc.) impactor (size range of 0.25-

497

0.5 µm, stage D) was used.12 A comparison of the Qsca, Qabs, and SSA of 500 nm diameter

498

particles for SRFA, ASOP, and heated ASOP is shown in Figure 10. ASOP have consistently

499

higher Qsca values than SRFA, with the highest scattering efficiency between 200-400 nm

500

decreasing rapidly with increasing wavelength. ASOP and SRFA particles have similar Qabs over

501

all wavelengths, with ASOP being slightly more absorbing. In both cases, Qabs decreases as the

502

wavelength increases. The absorbance enhancement at lower wavelengths is consistent with

503

previously reported optical properties for atmospheric BrC.15-16 The ASOP have a consistently

504

higher SSA than SRFA particles, and show a similar trend with a decreasing SSA as the

505

wavelength increases. The major difference between ASOP and SRFA is in the carbon content,

506

with the ASOP particles having a higher contribution of sp2 carbon and lower contributions of C-

507

OH and –COOH compared to SRFA particles. Additionally, minor contributions of Na and S in

508

SRFA may contribute to the differences in optical properties as well. The optical properties of

509

these particles are similar to other BrC species with an AAE of 1.8 for 200 nm size particles

510

which is in the range of 1.6-11 for BrC.15, 22 For the heated ASOP, the Qsca peaks at a wavelength

511

of approximately 780 nm, and heated ASOP consistently scatter more than non-heated ASOP at

512

all wavelengths above 400 nm. The Qabs of heated ASOP is fairly consistent over the range of

513

200-400 nm with a slight increase from 600-800 nm. This relatively flat absorbance over all

514

wavelengths is consistent with BC species.16 This is in contrast to ASOP and SRFA which are

28 ACS Paragon Plus Environment

Page 29 of 38

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

515

highly absorbing at lower wavelengths and decrease in absorptivity at higher wavelengths;

516

characteristic trends of BrC. Heated ASOP show invariable SSA across all wavelengths

517

consistent with the optical properties of carbonized materials. The effect of retrieving an accurate

518

RI over a wide wavelength range coupled with knowledge of the size distribution of particles

519

will help assess the effects of ASOP on climate forcing.

520

29 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

Page 30 of 38

521

Figure 10. Comparison of the a) the scattering efficiency (Qsca), b) absorption efficiency (Qabs),

522

and c) single scattering albedo (SSA) of 500 nm sized particles of SRFA, ASOP, and heated

523

ASOP particles.

524 525

526 527

Conclusions: The RI values for ASOP and SRFA particles were determined using STEM/EELS on a

528

single particle basis. SRFA is a common standard for retrieving optical properties and is similar

529

to ASOP since they are both from biological degradation products. The RISRFA was 1.22-0.07i at

530

532 nm, which is a slightly lower real part and higher imaginary part than retrieved through

531

optical methods. The RIASOP ranged from 1.21-0.11i to 1.29-0.08i over the 200-1200 nm

532

wavelength region. These values are similar to the retrieved values for amorphous carbon17 and

533

lower than tar balls40 by EELS. EELS retrieval of the RI of less absorbing particles (such as

534

SRFA and ASOP) overestimates the imaginary part of the RI for particles with a low absorption,

535

which in turn will lead to an underestimation of the real part of the refractive index. The optical

536

properties were related to particle compositions and the NEXAFS data indicated that ASOP had

537

higher fractions of sp2 carbon while the SRFA particles showed higher contributions from R-OH

538

and –COOH functional groups. The ASOP composition is consistent with ‘free-light’ SOM

539

indicating that the particles are likely from decomposition of natural organic materials. The

540

higher sp2 carbon fractions present in the ASOP results it an overall higher RI than that of SRFA.

541

When compared to previous literature with the EELS based retrievals of RI, the RIASOP values

542

were much lower and distinct from tar balls and were similar to values retrieved for amorphous

30 ACS Paragon Plus Environment

Page 31 of 38

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

543

carbon. Heating of the ASOP led to charring and removal of the oxygen-rich exterior of the

544

particles as shown by EELS and NEXAFS. Correspondingly, this led to higher values of

545

retrieved RI that were similar to those of graphitic carbon. Since the optical properties of BrC

546

species are not commonly categorized or used in atmospheric models, incorporation of the

547

RIASOP values for modeling predictions in the regions where they might be a considerable

548

fraction of the aerosol population can lead to better modeling of the radiative forcing. Proposed

549

mechanisms for formation of ASOP are through impaction of rain droplets on soil.28 It is

550

suggested that ASOP would be abundant in areas that have exposed open soils, such as

551

agricultural areas or grassland, that experience large amounts of rainfall.12 Various factors affect

552

the emission rates of ASOP: rainfall intensity, rain drop size, relative humidity and temperature

553

after rainfall, and seasonal variation that affect chemical composition of water soluble soil

554

organic matter. Additional studies are needed in order to understand the emission mechanism of

555

ASOP, contribution to the total aerosol budget, and variability in their chemical composition.

556

Suitable standards that have a similar refractive index to the particles of interest and are robust

557

under the electron beam are needed to calibrate the retrieved RI. Since the RIASOPh was very

558

close to the RI of graphitic carbon, the EELS method to retrieve the RI of single particles could

559

be more robust for highly absorbing species. For example, highly absorbing flame soot64 and

560

slightly absorbing humic-like substances65 can be used for further validation of the technique.

561

By retrieving the optical constants based on EELS measurements of individual particles within

562

multicomponent particles samples, radiative properties of individual components can be

563

evaluated for subsequent use in atmospheric models.

564

565

Supporting Information: 31 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

566 567

Back trajectories for field collected samples (Figure S1); Comparison of consecutive spectrum

568

for a single SRFA particle and STEM image after approximately 10 minutes of beam exposure

569

(Figure S2); Single scattering distributions for SRFA, ASOP, and heated ASOP (Figure S3);

570

Comparison of the RI of polystyrene spheres retrieved from EELS and CRD measurements

571

(Figure S4); Size distribution of ASOP (Figure S5); Tabulated data of ASOP (Table S1)

Page 32 of 38

572

573

Acknowledgements:

574

We are grateful to D. Bonanno and G. Kulkarni for assistance in sample collection at the SGP

575

site. We are grateful the ARM field site staff for assistance in sample collection at the Southern

576

Great Plains. The Pacific Northwest National Laboratory (PNNL) group acknowledges support

577

from the Science Acceleration Project of the Environmental Molecular Sciences Laboratory

578

(EMSL) at PNNL. The Lawrence Berkeley National Laboratory (LBNL) group acknowledges

579

support from the US Department of Energy's Atmospheric System Research Program, an Office

580

of Science, Office of Biological and Environmental Research (OBER). J. W. acknowledges the

581

student exchange program between the University of Würzburg and U.C. Berkeley (curator

582

Professor A. Forchel, Würzburg and NSF IGERT program at UCB, DGE-0333455, Nanoscale

583

Science and Engineering - From Building Blocks to Functional Systems.) The CCSEM and

584

STEM/EELS analyses were performed at EMSL, a National Scientific User Facility sponsored

585

by OBER at PNNL. PNNL is operated by the US Department of Energy by Battelle Memorial

586

Institute under contract DE-AC06-76RL0. STXM/NEXAFS analysis at beamlines 5.3.2.2 and

587

11.0.2.2 of the Advanced Light Source at LBNL is supported by the Director, Office of Science,

588

Office of Basic Energy Sciences of the US Department of Energy under Contract No. DE-AC0232 ACS Paragon Plus Environment

Page 33 of 38

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

589

05CH11231. We acknowledge use of the NOAA Air Resources Laboratory for the provision of

590

the HYSPLIT transport and dispersion model and READY website (http://www.ready.noaa.gov)

591

used in this publication.

592 593

33 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

594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639

Page 34 of 38

References: 1. Liu, L.; Mischchenko, M. I.; Arnott, W. P., A study of radiative properties of fractal soot aggregates using the superpostion T-matrix method. J. Quant. Spectrosc. RA 2008, 109 (15), 2656-2663. 2. China, S.; Scarnato, B.; Owen, R. C.; Zhang, B.; Ampadu, M. T.; Kumar, S.; Dzepina, K.; Dziobak, M. P.; Fialho, P.; Perlinger, J. A.; Hueber, J.; Helmig, D.; Mazzoleni, L. R.; Mazzoleni, C., Morphology and mixing state of aged soot particles at a remote marine free troposphere site: Implications for optical properties. Geophys. Res. Lett. 2015, 42 (4), 2014GL062404. 3. Kahnert, M.; Devasthale, A., Black carbon fractal morphology and short-wave radiative impact: A modelling study. Atmos. Chem. Phys. 2011, 11 (22), 11745-11759. 4. Veghte, D. P.; Moore, J. E.; Jensen, L.; Freedman, M. A., Influence of shape on the optical properties of hematite aerosol. J. Geophys. Res. - Atmos. 2015, 120 (14), 7025-7039. 5. Veghte, D. P.; Altaf, M. B.; Haines, J. D.; Freedman, M. A., Optical properties of non-absorbing mineral dust components and mixtures. Aerosol Science and Technology 2016, 50 (11), 1239-1252. 6. Boucher, O. D.; Randall, D.; Artaxo, P.; Bretherton, C.; Feingod, G.; Forster, P.; Kerminen, V. M.; Kondo, Y.; Liao, H.; Lohmann, U.; Rasch, P.; Satheesh, S. K.; Sherwood, S.; Stevens, B.; Zhang, X. Y., Coulds and Aerosols In: Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press: Cabridge, United Kingdom and New York, NY, USA, 2013. 7. Ramanathan, V.; Crutzen, P. J.; Kiehl, J. T.; Rosenfeld, D., Aerosols, Climate, and the Hydrological Cycle. Science 2001, 294 (5549), 2119-2124. 8. Chung, C. E.; Ramanathan, V.; Decremer, D., Observationally constrained estimates of carbonaceous aerosol radiative forcing. P. Natl. Acad. Sci. 2012, 109 (29), 11624-11629. 9. Feng, Y.; Ramanathan, V.; Kotamarthi, V. R., Brown carbon: a significant atmospheric absorber of solar radiation? Atmos. Chem. Phys. 2013, 13, 8607-8621. 10. Zhang, Y.; Forrister, H.; Liu, J.; Dibb, J.; Anderson, B.; Schwarz, J. P.; Perring, A. E.; Jimenez, J. L.; Campuzano-Jost, P.; Wang, Y.; Nenes, A.; Weber, R. J., Top-of-atmosphere radiative forcing affected by brown carbon in the upper troposphere. Nat. Geosci. 2017, 10, 486-489. 11. Zhu, J.; Crozier, P. A.; Anderson, J. R., Characterization of light-absorbing carbon particles at three altitudes in East Asian outflow by transmission electron microscopy. Atmos. Chem. Phys. 2013, 13, 6359-6371. 12. Wang, B.; Harder, T. H.; Kelly, S. T.; Piens, D. S.; China, S.; Kovarik, L.; Keiluweit, M.; Arey, B. W.; Gilles, M. K.; Laskin, A., Airborne Soil Organic paricles generated by precipitaition. Nat. Geosci. 2016, 9, 433-437. 13. Hand, J. L.; Malm, W. C.; Laskin, A.; Day, D.; Lee, T.; Wang, C.; Carrico, C.; Carrillo, J.; Cowin, J. P.; Collett, J.; Iedema, M. J., Optical, physical, and chemical properties of tar balls observed during the Yosemite Aerosol Characterization Study. J. Geophys. Res. 2005, 110 (D21210). 14. Hoffer, A.; Toth, A.; Nyiro-Kosa, I.; Pofai, M.; Gelencser, A., Light absorption properties of laboratory-generated tar ball particles. Atmos. Chem. Phys. 2016, 16, 239-246. 15. Laskin, A.; Laskin, J.; Nizkorodov, S. A., Chemistry of atmospheric brown carbon. Chem. Rev. 2015, 115 (10), 4335-4382. 16. Andreae, M.; Gelencser, A., Black carbon or brown carbon? The nature of light-absorbing carbonaceous aerosols. Atmos. Chem. Phys. 2006, 6 (10), 3131-3148. 17. Zhu, J.; Crozier, P. A.; Ercius, P.; Anderson, J. R., Derivation of Optical Properties of Carbonaceous Aerosols by Monochromated Electron Energy-Loss Spectroscopy. Microsc. Microanal. 2014, 20, 748-759.

34 ACS Paragon Plus Environment

Page 35 of 38

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

640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687

ACS Earth and Space Chemistry

18. Saleh, R.; Marks, M.; Heo, J.; Adams, P. J.; Donahue, N., M.; Robinson, A. L., Contribution of brown carbon and lensing to the direct radiative effect of carbonaceous aerosols form biomass burning. J. Geophys. Res. - Atmos. 2015, 120 (10), 285-10. 19. Lack, D.; Cappa, C. D., 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, 4207-4220. 20. Cappa, C. D.; Onasch, T. B.; Massoli, P.; Worsnop, D. R.; Bates, T. S.; Cross, E. S.; Davidovits, P.; Hakala, J.; Hayden, K. L.; Jobson, B. T.; Kolesar, K. R.; Lack, D.; Lerner, B. M.; Li, S.-M.; Mellon, D.; Nuaamann, I.; Olfert, J. S.; Petaja, T.; Quinn, P. K.; Song, C.; Subramanian, R.; Wiliams, E. J.; Zaveri, R. A., Radiative absorption enhancements due to the mixing state of atmospheric black carbon. Science 2012, 337 (6098), 1078-1081. 21. Liu, S.; Aiken, A. C.; Gorkowski, K.; Dubey, M. K.; Cappa, C. D.; Williams, L. R.; Herndon, S. C.; Massoli, P.; Fortner, E. C.; Chhabra, P. S.; Brooks, W. A.; Onasch, T. B.; Jayne, J. T.; Worsnop, D. R.; China, S.; Sharma, N.; Mazzoleni, C.; Xu, L.; Ng, N. L.; Liu, D.; Allan, J. D.; Lee, J. D.; Fleming, Z. L.; Mohr, C.; Zotter, P.; Szidat, S.; Prevot, A., Enhanced light absorption by mixed source black and brown carbon particles in UK winter. Nat. Commun. 2015, 6. 22. Moise, T.; Flores, J. M.; Rudich, Y., Optical properties of secondary organic aerosols and their changes by chemical processes. Chem. Rev. 2015, 115 (10), 4400-4439. 23. Lack, D.; Langridge, J., On the attribution of black and brown carbon light absorption using the Angstrom exponent. Atmos. Chem. Phys. 2013, 13 (20), 10535-10543. 24. Bond, T. C.; Doherty, S. J.; Fahey, D. W.; Forster, P.; Berntsen, T.; DeAngelo, B. J.; Flanner, M. G.; Ghan, S.; Karcher, B.; Koch, D.; Kinne, S.; Kondo, Y.; Quinn, P. K.; Sarofim, M. C.; Schultz, M. G.; M., S.; Venkataraman, C.; Zhang, H.; Zhang, S.; Bellouin, N.; Guttikunda, S. K.; Hopke, P. K.; Jacobson, M. Z.; Kaiser, J. W.; Z., K.; Lohmann, U.; Schwarz, J. P.; Shindell, D.; Storelvmo, T.; Warren, S. G.; Zender, C. S., Bounding the role of black carbonin the climate system: A scientific assessment. J. Geophys. Res. Atmos. 2013, 118, 5380-5552. 25. Lewis, K.; Arnott, W. P.; Moosmuller, H.; Wold, C. E., Strong spectral variation of biomass smoke light absorption and single scatterin albedo observed with a novel dual-wavelength photoacoustic instrument. J. Geophys. Res. - Atmos. 2008, 113 (D16). 26. Graber, E. R.; Rudich, Y., Atmospheric HULIS: How humic-lie are they? A comprehensive and critical review. Atmos. Chem. Phys. 2006, 6, 729-753. 27. Cotrufo, M. F.; Soong, J. L.; Horton, A. J.; Campbell, E. E.; Haddix, M. L.; Wall, D. H.; Parton, W. J., Formation of soil organic matter via biochemical and physical pathways of litter mass loss. Nat. Geosci. 2015, 8, 776-779. 28. Joung, Y. S.; Bruie, C. R., Aerosol Generation by Raindrop Impact on Soil. Nat. Commun. 2015, 6, 6083. 29. Lang-Yona, N.; Rudich, Y.; Segre, E.; Dinar, E.; Abo-Riziq, A., Complex refractive indices of aerosols retrieved by continuous wave-cavity ring down aerosol spectrometer. Analytical Chemistry 2009, 81 (5), 1762-1769. 30. Dinar, E.; 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. 2008, 137, 279-295. 31. Washenfelder, R. A.; Flores, J. M.; Brock, C. A.; Brown, S. S.; Rudich, Y., Broadband measurements of aerosol extinction in the ultraviolet spectral region. Atmos. Meas. Tech. 2013, 6, 861877. 32. Flores, J. M.; Washenfelder, R. A.; Adler, G.; Lee, H. J.; Segev, L.; Laskin, J.; Laskin, A.; Nizkorodov, S. A.; Brown, S. S.; Rudich, Y., Complex refractive indices in the near-ultraviolet spectral region of biogenic secondary organic aerosol aged with ammonia. Phys. Chem. Chem. Phys. 2014, 16. 35 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

688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735

Page 36 of 38

33. Hudson, P. K.; Gibson, E. R.; Young, M. A.; Kleiber, P. D.; Grassian, V. H., A newly designed and constructed instrument for coupled infrared extinction and size distribution measurements of aerosols. Aerosol Sci. Technol. 2007, 41 (7), 701-710. 34. Moosmuller, H.; Chakrabary, R. K.; Arnott, W. P., Aerosol light absorption and its measurement: A review. J. Quant. Spectrosc. RA 2009, 110 (11), 844-878. 35. Baynard, T.; Lovejoy, E. R.; Pettersson, A.; Brown, S. S.; Lack, D.; Osthoff, H.; Massoli, P.; Ciciora, S.; Dube, W. P.; Ravishankara, A. R., Design and application of a pulsed cavity ring-down aerosol extinction spectrometer for field measuremtns. Aerosol Sci. Technol. 2007, 41, 447-462. 36. Lack, D.; Lovejoy, E. R.; Baynard, T.; Pettersson, A.; Ravishankara, A. R., Aerosol absorption measurement using photoacoustic spectroscopy: sensitivity, calibration, and uncertainty develoopments. Aerosol Sci. Technol. 2006, 40, 697-708. 37. Lack, D.; Moosmuller, H.; McMeeking, G. R.; Chakrabary, R. K.; Baumgardner, D., Characterizing elemental, equivalent black, and refractory black carbon aerosol particles: a review of techniques, their limitations and uncertainties. Anal. Bioanal. Chem. 2014, 406, 99-122. 38. Ebert, M.; Winbruch, S.; Rausch, A.; Gorzawski, G.; Helas, G.; Hoffman, P.; Wex, H., The complex refractive index of aerosols during LACE 98 as derived from the analysis of individual particles. J. Geophys. Res. Atmos. 2002, 107 (D21), 8121. 39. Ebert, M.; Weinbruch, S.; Hoffman, P.; Ortner, H. M., The chemical composition and complex refractive index of rural and urban influenced aerosols determined by individual particle analysis. Atmos. Environ. 2004, 38 (38), 6531-6545. 40. 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. 41. Bluvshtein, N.; Flores, J. M.; Segev, L.; Rudich, Y., A new approach for retrieving the UV-vis optical properties of ambient aerosols. Atmos. Meas. Tech. 2016, 9, 3477-3490. 42. Draxler, R. R.; D., R. G., HYSPLIT Model. NOAA ARL READY, 2012. 43. Laskin, A.; Cowin, J. P.; Iedema, M. J., Analysis of individual environmental particles usig modern methods of electron microscopy and x-ray microanalysis. J. Electron Spectrosc. Relat. Phenom. 2006, 150, 260-274. 44. Laskin, A., Electron beam analysis and microscopy of individual particles. In Fundamentals and Applications in Aerosol Spectroscopy. Taylor and Francis Books: New York, 2010. 45. Moffet, R. C.; Henn, T. R.; Tivanski, A. V.; Hopkins, R. J.; Desyaterik, Y.; Kilcoyne, A. L. D.; Tyliszczak, T.; Fast, J.; Barnard, J.; Shutthanandan, V.; Cliff, S. S.; Perry, K. D.; Laskin, A.; Gilles, M. K., Microscopic characterization of carbonaceous aerosol particle aging in the outflow from Mexico City. Atmos. Chem. Phys. 2010, 10, 961-976. 46. Moffet, R. C.; Tivanski, A. V.; Gilles, M. K., Scanning x-ray transmission microscopy: applications in atmospheric microscopy. In Fundamentals and Applications in Aerosol Spectroscopy. Taylor and Francis Books: New York, 2010. 47. Hopkins, R. J.; Tivanski, A. V.; Marten, B. D.; Gilles, M. K., Chemical bonding and structure of black carbon reference materials and individual carbonaceous atmospheric aerosols. J. Aerosol Sci. 2007, 38, 573-591. 48. Stoger-Pollach, M., Optical properties and bandgaps from low loss EELS: Pitfalls and solutions. Micron 2008, 39, 1092-1110. 49. Koch, E. E.; Otto, A., Characteristic energy losses of 30 keV - electrons in vapours of aromatic hydrocarbons. Optics Comm. 1969, 1 (2), 47-49. 50. Ritchie, R. H., Plasma Losses by Fast Electrons in Thin Films. Phys. Rev. 1957, 106 (5), 874-881. 51. Egerton, R. F., Electron Energy-Loss Spectroscopy in the Electron Microscop. 3rd ed.; Springer: New York, Dordrecht, Heidelber, London, 2001. 52. Mätzler, C. IAP research Report No. 2002-08; Universität Bern: Bern, Switzerland, 2002. 36 ACS Paragon Plus Environment

Page 37 of 38

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

736 737 738 739 740 741 742 743 744 745 746 747

53. O'Brien, R. E.; Neu, A.; Epstein, S. A.; MacMillan, A. C.; Wang, B.; Kelly, S. T.; Nizkorodov, S. A.; Laskin, A.; Moffet, R. C.; Gilles, M. K., Physical properties of ambient and laboratory-generated secondary organic aerosol. Geophys. Res. Lett. 2014, 41 (12), 4347-4353. 54. Yu, J. Z.; Xu, J.; Yang, H., Charring characteristics of atmospheric organic particulate matter in thermal analysis. Environ. Sci. Technol. 2002, 36, 754-761. 55. Keiluweit, M.; Nico, P. S.; Johnson, M. G.; Kleber, M., Dynamic molecular structure of plant biomass-derived black carbon (biochar). Environ. Sci. Technol. 2010, 44, 1247-1253. 56. Maria, S. F.; Russell, L. M.; Gilles, M. K.; Myneni, S. C., Organic aerosol growth mechanisms and their climate-forcing implications. Science 2004, 306, 1921-1924. 57. Tivanski, A. V.; Hopkins, R. J.; Tyliszczak, T.; Gilles, M. K., Oxygenated interface on biomass burn tar balls determined by single particle scanning transmission x-ray microscopy. J. Phys. Chem. A 2007, 111, 5448-5458

748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768

58. Solomon, D.; Lehmann, J.; Kinyangi, J.; Liang, B.; Schafer, T., Carbon k-edge NEXAFS and FTIR-ATR spectroscopic investigation of organic carbon speciation in soils. Soil Sci. Soc. Am. J. 2005, 69, 107-119. 59. Bernard, S.; Beyssac, O.; Benzerara, K.; Findling, N.; Tzvetkov, G.; Brown, G. E., XANES, Raman, and XRD study of anthracene-based cokes and saccharose-based chars submitted to high-temperature pyrolysis. Carbon 2010, 48 (9), 2506-2516. 60. Matheson, L. A.; Saunderson, J. L., Optical and Elctrical Properties of Polystyrene. In Sytrene: Its Polymers, Copolymers and Derivatives. Hafner Publishing Corporation: Darian, Connecticut, 1970. 61. Posfai, Atmospheric tar balls: Particles from biomass and biofuel burning. J. Geophys. Res. 2004, 109 (D06213). 62. Chakrabary, R. K.; Moosmuller, H.; Chen, L. W. A.; Lewis, K.; Arnott, W. P.; Mazzoleni, C.; Dubey, M. K.; Wold, C. E.; Hao, W. M.; Kreidenweis, S. M., Brown carbon in tar balls from smoldering biomass combustion. Atmos. Chem. Phys. 2010, 10, 6363-6370. 63. Hopkins, R. J.; Lewis, K.; Desyaterik, Y.; Wang, Z.; Tivanski, A. V.; Arnott, W. P.; Laskin, A.; Gilles, M. K., Correlations between optical, chemical, and physical properties of biomass burn aerosols. Geophys. Res. Lett. 2007, 34, L18806. 64. Chang, H.; Charalampopoulos, T. T., Determination of the wavelength dependence of refractive indices of flame soot. Proc. R.Soc. London 1990, 1, 577-591. 65. Hoffer, A.; Gelencser, A.; Guyon, P.; Kiss, G.; Schmid, O.; Frank, G. P.; Artaxo, P.; Andreae, M., Optical properties of humic-like substances (HULIS) in biomass-burning aerosols. Atmos. Chem. Phys. 2006, 6, 3563-3570.

769

37 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

247x85mm (72 x 72 DPI)

ACS Paragon Plus Environment

Page 38 of 38