Atmospheric Reactivity of Fullerene (C60) Aerosols - ACS Earth and

Nov 30, 2017 - Cloud seeding, where an increase in polarity on the surface of the nanoparticles may cause them to act as more efficient cloud condensa...
0 downloads 7 Views 4MB Size
Subscriber access provided by READING UNIV

Article 60

Atmospheric Reactivity of Fullerene (C ) Aerosols Dhruv Mitroo, Jiewei Wu, Peter F Colletti, Seung Soo Lee, Michael Walker, William H. Brune, Brent J. Williams, and John D. Fortner ACS Earth Space Chem., Just Accepted Manuscript • DOI: 10.1021/ acsearthspacechem.7b00116 • Publication Date (Web): 30 Nov 2017 Downloaded from http://pubs.acs.org on December 9, 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 26 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

ACS Earth and Space Chemistry

Title

2 3

Atmospheric Reactivity of Fullerene (C60) Aerosols

4 5

Author List

6 7

Dhruv Mitroo,†,‡ Jiewei Wu,† Peter F. Colletti,† Seung Soo Lee,† Michael J. Walker,† William H.

8

Brune,§ Brent J. Williams,*,† and John D. Fortner,*,†

9 10



11

Louis, St. Louis, Missouri 63130, United States

Department of Energy, Environmental and Chemical Engineering, Washington University in St.

12 13

§

14

University Park, Pennsylvania 16802, United States

Department of Meteorology and Atmospheric Sciences, The Pennsylvania State University,

15 16

Table of Contents Graphic

17 18 19 20

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

21

Abstract

22 23

Rapid growth and adoption of nanomaterial-based technologies underpins a risk for unaccounted

24

material release to the environment. Carbon based materials, in particular fullerenes, have been

25

widely proposed for a variety of applications. A quantitative understanding of how they behave

26

is critical for accurate environmental impact assessment. While their aqueous phase reactivity,

27

fate, and transport has been studied for over a decade, aerosol phase reactivity remains

28

unexplored. Here, the transformation of C60, as nanocrystal (nC60) aerosols, is evaluated over a

29

range of simulated atmospheric conditions. Upon exposure to UV light, gas-phase O3, and ·OH,

30

nC60 is readily oxidized. This reaction pathway is likely limited by diffusion of oxidants

31

within/through the nC60 aerosol. Further, gas-phase oxidation induces disorder in the crystal

32

structure without affecting aerosol (aggregate) size. Loss of crystallinity suggests aged nC60

33

aerosols will be less effective ice nuclei; but, an increase in surface oxidation will improve their

34

cloud condensation nuclei ability.

35 36 37 38 39 40 41 42

Keywords: fullerenes, nC60, nanocrystals, oxidation, potential aerosol mass (PAM) reactor,

43

atmospheric aging

2 ACS Paragon Plus Environment

Page 2 of 26

Page 3 of 26 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

44

ACS Earth and Space Chemistry

Introduction

45 46

Industrial scale production and adoption of engineered nanomaterials raises concerns over their

47

inadvertent impacts on human health and natural systems.1,2 Of these, carbon based nanoscale

48

materials such as fullerenes, carbon nanotubes (CNTs), and graphene are being developed for a

49

variety of applications.3–5 Fundamental understanding of how such materials behave when

50

released in fluvial, marine, or atmospheric systems is critical for material design, life cycle

51

analyses, risk assessment, and eventual effective waste management scenarios. In contrast to

52

aqueous systems, atmospheric aerosol release scenarios have only recently been considered for

53

engineered nanomaterials, such as carbon nanotubes (CNTs) and fullerenes, and inorganic

54

nanoparticles, such as TiO2.2,6,7 In particular, the fate of fullerenes in aqueous systems is being

55

studied extensively. 8–10 In the atmosphere, fullerenes have been identified in sooty flames11 and

56

lightning12: high energy (and high temperature) processes. To date, there is little understanding

57

of how low energy events (relative to formation events) such as atmospheric processing (e.g.,

58

photo-oxidation)13 of fullerenes significantly alter their physicochemical properties – and thus

59

their behavior, which is a critical knowledge gap for accurate modeling of solubility,

60

(bio)availability, reactivity, stability, and toxicity.

61

In the aqueous phase, C60 (as nC60) is readily oxidized in absence of light with both

62

ozone, yielding hemiketal and hydroxyl functionalities proposed via a modified Criegee

63

mechanism, and by hydroxyl radicals, forming hydroxylated fullerenes.8,14 Recently, Wu et al.

64

found that free chlorine also oxidizes nC60 in aqueous media, eventually altering the

65

nanoparticles’ crystal structure.15 Because the cage structure imposes strain on its aromatic rings,

66

epoxide formation is unstable, and water hydrolyzes intermediate C-O-O● structures to yield

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

67

C=O or C-OH functional groups.16–18 Further, UV light has been shown to sensitize aqueous

68

phase reactions leading to oxidation products including epoxides and ethers.19 A decade before

69

this discovery, Foote et al. proposed an energy transfer mechanism whereby photons excite 1C60

70

to an excited state, 1C60*, leading through intersystem crossing (ISC) to the triplet excited state

71

3

72

oxygen 3O2 to singlet oxygen, 1O2.20 Krusic et al. instead proposed a type I mechanism where

73

3

74

shown to generate reactive oxygen species (ROS).22 The generation of the radical anion has been

75

also observed in aqueous phase nC60 aggregates as well, also generating ROS.23

C60*.20 Then, through a type II energy transfer mechanism, this can convert excited triplet

C60* forms the anion C60-● in the presence of an electron donor.21 Both pathways have been

76

With regard to the atmosphere (troposphere), nC60 chemistry is largely unexplored. To

77

date there are two reported field observations of airborne fullerenes (without distinguishing

78

between their molecular and aggregated states),24,25 one systematic laboratory study of

79

atmospheric ozonolysis of fullerene,26 and one of atmospheric ozonolysis / photooxidation of

80

CNTs.27 Utsunomiya et al. observed fullerenes in samples from coal-fired power plants collected

81

in Detroit, MI,24 and Sanchis et al. report observations of aerosol-bound fullerene in all samples

82

collected in two campaigns covering different regions of the Mediterranean.25 It is worth noting

83

that in addition to the two campaigns, fullerenes have been identified in other airborne samples,28

84

further highlighting the importance of constraining fullerene mobilization in the atmosphere.

85

However, atmospheric processing of these fullerenes has not been investigated or discussed.

86

Recently, Tiwari et al. explored atmospheric ozonolysis of C60 bulk fullerene powder through lab

87

studies, showing that atmospheric timescales can accommodate surface reaction of potential

88

fullerene aerosols.26 Similarly, Liu and coworkers investigated atmospheric photo-oxidation of

89

CNTs via both ozone and hydroxyl attack, suggesting evidence of processing, and raising

4 ACS Paragon Plus Environment

Page 4 of 26

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

ACS Earth and Space Chemistry

90

concern over gas-particle partitioning of organics in CNTs during atmospheric evolution.27 In

91

both studies, commercially available C60 powder and CNTs of fixed (commercial grade) purity

92

and aggregate size were used. Other studies concerning aerosol-phase heterogeneous (gas-phase

93

reaction on the surface of a solid particle) oxidation via ozone or hydroxyl radicals have

94

investigated the chemistry of non-aromatics such as oleic / stearic acid particles,29 squalene,30

95

palmitic acid,31 and erythritol and levoglucosan;32 aromatics including surface-bound polycyclic

96

aromatic hydrocarbons (PAHs).33–35 To date, to the best of our knowledge, none have looked at

97

engineered nanoparticles.

98

Here, we present direct evidence of atmospherically relevant heterogeneous reactions of

99

nC60 aerosols by simultaneous ozone and hydroxyl radical oxidation in presence of UV light

100

using a Potential Aerosol Mass (PAM) reactor,36 an oxidative flow-tube reactor (OFR).

101

Atmospheric exposure, or mimicked ‘age’, of the aerosol spans between ~ 0 – 2 weeks, over

102

which, we monitor physical and chemical changes to the nC60 aerosols. Surface oxidation, crystal

103

structure, and aggregate (aerosol) size are described as a function of exposure in the gas-phase.

104 105

Materials and Methods

106 107

Setup. For this study, nC60 was synthesized as described previously in detail by Fortner et al.8

108

The experimental setup schematic for the aerosol-phase oxidation study is available in the

109

Supporting Information (Figure S1), alongside a description of the nC60 synthesis (section titled

110

nC60 nanocrystal synthesis). Briefly, a 25mgL-1 nC60 suspension was atomized in a 6-jet Collison

111

Nebulizer (BGI, Inc; Waltham, MA) and passed through an inlet heated to ~100°C to drive

112

surface water off the aerosolized particles. Half of the flow (corresponding to 5.0 L min-1) was

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

113

sent to the PAM reactor via a custom-built diffusion dryer to remove water vapor from the

114

aerosol stream. Relative humidity (RH) control in the PAM chamber was achieved by adding

115

4.6 L min-1 of N2 flow with fixed humidity. Additional details can be found in the Setup details

116

section in the Supporting Information. Reported negative ζ-potential measurements in nC60

117

aqueous suspensions37 suggest that electrostatic repulsion is important in stabilizing the

118

suspensions. If aggregation of primary nC60 nanoparticles does not occur in suspension, then

119

drying should not induce further clustering. This was the case for our system (see Supporting

120

Information Figure S2), where the mode of the size distribution in the aerosol phase was

121

comparable to that of the aqueous phase.

122

Reactor. The custom built PAM reactor is an iridite-treated aluminum cylinder, 18

123

inches in length and 8 inches in inner diameter, giving it a total volume of 13 L. The chamber is

124

equipped with two 12 inch mercury lamps with peak wavelengths at 185 nm and 254 nm (BHK

125

Inc. Analamp Model No. 82-9304-03; Ontario, CA) housed in Teflon sheaths. There is an

126

annular flow of N2 through the sheaths to prevent direct contact with the lamps and purge any

127

outgas products. The mercury lamps are connected to a variable voltage supply (BHK Inc.;

128

Ontario, CA) to regulate light intensity. The former wavelength is necessary for the O2

129

photolysis to produce O3, and the latter dissociates O3 to form singlet oxygen, O(1D), which in

130

the presence of H2O forms the hydroxyl radical.36 Teflon sheaths are transparent to both

131

wavelengths. Additional ozone is delivered to the inlet of the PAM to accelerate production of

132

hydroxyl radicals. A simplistic mechanism of oxidant formation can be found in the Supporting

133

Information, equations (3) – (7) and more refined mechanisms are available in the literature.38–42

134

At a fixed 10 L min-1 flowrate, the mean residence time for the nC60 aerosol in the reactor, τres,

135

(defined by the volume of the reactor divided by the outlet flowrate) is approximately 78 s. This

6 ACS Paragon Plus Environment

Page 6 of 26

Page 7 of 26 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

136

parameter is held constant throughout the study. Typically, heterogeneous reactions in OFRs

137

conducted by other groups are performed at RH in the range of ~10-60%.39,40,43,44 For our study,

138

RH was chosen at ~30% (dry conditions to avoid potential deliquescence of aerosols, yet high

139

enough to promote ·OH formation) and kept constant since the mixing ratio of water directly

140

affects hydroxyl radical production (as per Supporting Information eq. (7)).40

141

OH Exposure. The reactor can generate hydroxyl radicals in mixing ratios >1000 ppt,

142

which is several order-of-magnitudes higher than atmospheric concentrations (typically 10-19 cm2s-1, typical for crystalline

233

solids60), which is considerably longer than atmospheric reaction timescales. However, if

234

aerosol-phase surface reaction morphs the pristine structure, the diffusion timescales can be

235

greatly affected, and the observed rate of reaction would also change. For example, trace levels

236

of hydrogen chloride (HCl) gas have shown to induce disorder on ice surfaces in a temperature

237

range of 186 K to 243 K.61 This process explains why reported values for diffusion coefficients

238

of HCl on ice vary dramatically on different ice surfaces.62,63 If atmospheric photo-oxidants

239

could have similar effects on nC60 surfaces, reaction rates would not be as slow as expected. To

240

explore this, high resolution transmission electron microscopy (HR-TEM) images (Figure 5) and

241

X-ray diffraction (XRD) patterns (Figure S4) were evaluated for parent and oxidized samples.

242

Results suggest that crystallinity is lost upon oxidation; however, persistence of peaks at ca. 520

243

cm-1 and 570 cm-1 (parent peaks) in FTIR spectra suggest that parent C60 does exist, to some

244

degree in the aggregate (likely in the aerosol core). While this would challenge the suggestion

245

that reaction (chemical or physical) is surface-limited, without further investigation, it is not

246

straightforward to assess the underlying mechanism that drives the observed disorder.

247

Underpinning this process, several phenomena may be responsible, including: i) increased

248

surface disorder, either allowing oxidants to diffuse to the core more readily, or ii) rearrangement

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

249

of product molecules to the core that, although involving an enthalpic penalty, are entropically

250

preferred. Owing to the cage structure of a C60 molecule and its (caged) products, the enthalpic

251

cost of breaking a π−π stacking interactions is likely not strong enough to hinder disorder in the

252

nanoparticle. Other possibilities include iii) intercolated water within the nanoparticle aggregate,

253

which upon UV light irradiation, releases ROS that readily react with the core, or unexpected

254

ligand-induced rearrangements64 or iv) a purely UV-light sensitized rearrangement of the

255

nanoparticle structure that, although interesting, is not relevant to tropospheric chemistry. We

256

suggest excluding possibility ii) due to the unlikeliness of entropy outweighing enthalpy and iii)

257

due to FTIR data not showing dramatic decrease in the parent peaks upon oxidation. The only

258

way possibility iv) would be excluded is by assuming that work done by Lee et al.19, implying

259

that the free oxygen’s role has a fundamental effect on crystal structure too. OFRs like the PAM

260

reactor operate on τres values on the order of seconds to minutes, which is short compared to an

261

average aerosol lifetime in the troposphere. Therefore we can not quantify the effect of diffusion

262

on atmospheric photo-oxidation for nC60. David et al. suggest that >260 K, C60 molecules

263

undergo (continuous) reorientation;65 therefore, although atmospheric photooxidation is not a

264

high energy process, it may be enough to destabilize π – π stacking in the outer lattice and

265

induce disorder in the crystal structure.

266

For reacted samples, XPS spectra are different from those observed for samples produced

267

by the aqueous phase reaction of nC60 with ozone, as published by Fortner et al.8 Ozonation of

268

aqueous nC60 yields more di-oxidized than mono-oxidized carbon (30% and 18%, respectively,

269

from XPS fitted peak area), whereas in this study, fullerenes subject to the highest ·OH exposure

270

of ~2.24E+12 molec. cm-3 s (~17.3 equivalent days of aging) yielded comparatively more mono-

271

oxidized than di-oxidized carbon (27.8% and 12.3%, respectively, from XPS peak area), and the

12 ACS Paragon Plus Environment

Page 12 of 26

Page 13 of 26 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

272

remaining un-derivatized carbon fraction was higher (52% for the aqueous study and 60.0% for

273

this study). For these, resulting apectra are more similar to that of the aqueous phase reaction of

274

nC60 with ·OH as described by Lee et al., although the mono-oxidized carbon fraction of the area

275

is much higher for aqueous-phase oxidation (53.64% for the aqueous study and 27.8% for this

276

study at ~17.3 equivalent days of aging).

277

Environmental Implications. In this report, we observe direct evidence of atmospheric

278

oxidation of C60. However, in relatively dry conditions (~30% RH), even a very strong oxidizing

279

atmosphere yields a limited reaction of the material. Based on the decay of un-derivatized carbon

280

via XPS analysis, we estimate the atmospheric photochemical lifetime of nC60 to be 1-2 months,

281

meaning wet and dry deposition would remain the dominant atmospheric loss pathways. Such a

282

slow oxidation mechanism helps explain detection of aerosol-bound fullerenes even in the

283

Mediterranean basin,25 far removed from anthropogenic sources. The reactive and effective

284

uptake coefficients for nC60 to O3 or ·OH has not been calculated due to coupling of the oxidants

285

(hydroxylation photochemistry versus ozonolysis) as well as complex mixing patterns in the

286

reactor, and while important, are beyond the scope of this work. It is worth mentioning that Pillar

287

et al.66 assessed that, using an effective uptake coefficient of ozone on catechol at the air-water

288

interface of 2.73E-7 and an effective uptake of ·OH between 0.01 and 1, the rate of reaction of

289

O3 to catechol can be 1 – 100 times that of ·OH to catechol for mean tropospheric values.

290

However, this is dependent on forming the initial ozonide, which given C60’s caged structure,

291

becomes less favorable. Given our estimate of nC60’s photochemical lifetime is on the order of

292

months, compared to hours for catechol from Pillar et al., we take caution in extending their

293

values to our system. Taken together, with light sensitizing the reaction by both producing ·OH

294

and promoting C60 to its excited triplet state, we believe both in the PAM and in the troposphere,

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

295

that ·OH oxidation accounts for more nC60 depletion than O3 oxidation. We do however

296

recognize that in the PAM this reaction could be augmented artificially due to the UV light

297

required in ·OH generation.

298

Of interest in future work is to understand interactions with fog water or cloud water

299

(e.g., how uptake kinetics vary as a function of dry aerosol age) using both dry aerosolization

300

techniques (Tiwari et al.) and wet aerosolization techniques (this work). Functionalizing the

301

surface should increase the hydrophilicity of the nanoparticle, however if, once scavenged, the

302

outer layer sheds (is solubilized) into solution, the freshly exposed hydrophobic surface may

303

repartition to the aerosol phase, where it can undergo further aerosol phase oxidation. This could

304

be important for reactions in deliquesced aerosols. Cloud seeding, where an increase in polarity

305

on the surface of the nanoparticles may cause them to act as more efficient cloud condensation

306

nuclei (CCN), can become important at tropical latitudes where clouds are warmer, but not as

307

effective at higher latitudes or altitudes where ice nuclei (IN) become more important.

308

Furthermore, wildfires and agricultural activity at those latitudes result in a consistent source of

309

soot aerosols, likely containing fullerene. Increased light exposure yields consistent vertical

310

profiles of ·OH45 which can become a dominant oxidant at higher altitudes, resulting in

311

hydroxylation photochemistry driving fullerene reactivity in the free troposphere.

312 313

14 ACS Paragon Plus Environment

Page 14 of 26

Page 15 of 26 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

314 315

Figure 1: Change in the geometric mean mobility diameter (Dpg) of aerosolized nC60 as a

316

function of OH Exposure for mixed OH / O3 oxidation. Values are normalized by an average Dpg

317

for unoxidized nC60 (control, see SI), and the bias associated with the control Dpg is subtracted.

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

318 319

Figure 2: Offset FTIR spectra for nC60 samples collected on PTFE filters at different levels of

320

OH Exposures for mixed OH / O3 oxidation. Peaks around ca. 1400 cm-1 and 1700 cm-1 indicate

321

of C – OH and C = O groups respectively.

322 323 324 16 ACS Paragon Plus Environment

Page 16 of 26

Page 17 of 26 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

325 326

Figure 3: XPS data on nC60 for a) aqueous sample and b) aerosol sample without exposure, and

327

for aerosol nC60 exposed at c) 2.4 d) 7.9 e) 13.3 and f) 17.2 equivalent days of aging. Peaks 1, 2,

328

and 3 correspond to underivatized, mono-oxidized, and di-oxidized carbon in that order.

329

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

330 331

Figure 4: Integrated peak areas from XPS data as a function of OH Exposure for differently

332

hybridized carbon.

333

18 ACS Paragon Plus Environment

Page 18 of 26

Page 19 of 26 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

334 335

Figure 5: Transmission electron microscopy data for unoxidized nC60 displayed as a) HR-TEM

336

image and b) fast Fourier Transform (FFT) pattern, and oxidized nC60 (16.1 equivalent days of

337

aging) displayed as c) HR-TEM image and d) FFT pattern.

338 339

TABLES

-3

OH Exposure (molec. cm s) (3.14±0.16) E+11 (1.02±0.01) E+12 (1.73±0.02) E+12 (2.24±0.03) E+12

Equivalent Age (days)

0 2.42±0.12 7.87±0.10 13.3±0.1 17.2±0.20

Underivatized Carbon % Area 80.00 72.89 68.93 62.12 60.00

19 ACS Paragon Plus Environment

Monooxidized Carbon % Area 16.29 22.61 25.05 28.66 27.75

Di-oxidized Carbon % Area 3.72 4.50 6.03 9.22 12.25

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

340 341

Table 1: XPS Peak fitted data.

342 343

Associated Content

344 345

Supporting Information

346

Experimental details and additional data.

347 348

Author Information

349 350

Corresponding Authors

351

*Address: One Brookings Drive, Brauer 1015, St. Louis, Missouri, 63130. Email (J.D.F.):

352

[email protected]. Telephone: (314) 935-9293.

353

*Address: One Brookings Drive, Brauer 3026, St. Louis, Missouri, 63130. Email (B.J.W.):

354

[email protected]. Telephone: (314) 935-9279.

355

Present Addresses

356



357

Miami, Florida, 33149, USA.

358

Notes

359

The authors declare no competing financial interest.

Dhruv Mitroo: Rosenstiel School of Marine and Atmospheric Sciences, University of Miami,

360 361

Acknowledgments

362

20 ACS Paragon Plus Environment

Page 20 of 26

Page 21 of 26 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

363

This work was supported by the National Science Foundation Grant #1236865. We would like to

364

thank Adan Montoya, Carl Hinton, and Yining Ou for their help in washing nC60 batches. We

365

would also like to thank Dr. Huafang Li and the Institute of Materials Science & Engineering

366

(IMSE) at Washington University in St. Louis (WUStL) for her valuable assistance and training

367

on the XPS. Finally, we sincerely thank Prof. James Ballard at WUStL and Prof. Dylan Millet at

368

the University of Minnesota (UMN) for insightful discussion and valuable feedback.

369 370

References

371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397

(1)

Nowack, B. The behavior and effects of nanoparticles in the environment. Environ. Pollut. 2009, 157 (4), 1063–1064 DOI: 10.1016/j.envpol.2008.12.019. (2) Biswas, P.; Wu, C.-Y. Nanoparticles and the Environment. J. Air Waste Manag. Assoc. 2005, 55 (6), 708–746 DOI: 10.1080/10473289.2005.10464656. (3) Wong, S.; Karn, B. Ensuring sustainability with green nanotechnology. Nanotechnology 2012, 23 (29), 290201–290201 DOI: 10.1088/0957-4484/23/29/290201. (4) Bakry, R.; Vallant, R. M.; Najam-ul-Haq, M.; Rainer, M.; Szabo, Z.; Huck, C. W.; Bonn, G. K. Medicinal applications of fullerenes. Int. J. Nanomedicine 2007, 2 (4), 639–649. (5) Tremblay, Jean-Francois. Mitsubishi Chemical Aims at Breakthrough. Chem. Eng. News 2002, 80 (49), 16–17 DOI: 10.1021/cen-v080n049.p016. (6) Buseck, P. R.; Adachi, K. Nanoparticles in the Atmosphere. Elements 2008, 4 (6), 389– 394 DOI: 10.2113/gselements.4.6.389. (7) Nowack, B.; Bucheli, T. Occurrence, behavior and effects of nanoparticles in the environment. Environ. Pollut. 2007, 150 (1), 5–22 DOI: 10.1016/j.envpol.2007.06.006. (8) Fortner, J. D.; Kim, D.-I.; Boyd, A. M.; Falkner, J. C.; Moran, S.; Colvin, V. L.; Hughes, J. B.; Kim, J.-H. Reaction of Water-Stable C 60 Aggregates with Ozone. Environ. Sci. Technol. 2007, 41 (21), 7497–7502 DOI: 10.1021/es0708058. (9) Li, Y.; Niu, J.; Shang, E.; Crittenden, J. C. Synergistic Photogeneration of Reactive Oxygen Species by Dissolved Organic Matter and C 60 in Aqueous Phase. Environ. Sci. Technol. 2015, 49 (2), 965–973 DOI: 10.1021/es505089e. (10) Sayes, C. M.; Fortner, J. D.; Guo, W.; Lyon, D.; Boyd, A. M.; Ausman, K. D.; Tao, Y. J.; Sitharaman, B.; Wilson, L. J.; Hughes, J. B.; et al. The Differential Cytotoxicity of WaterSoluble Fullerenes. Nano Lett. 2004, 4 (10), 1881–1887 DOI: 10.1021/nl0489586. (11) Howard, J. B.; McKinnon, J. T.; Makarovsky, Y.; Lafleur, A. L.; Johnson, M. E. Fullerenes C60 and C70 in flames. Nature 1991, 352 (6331), 139–141 DOI: 10.1038/352139a0.

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

398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442

(12) Heymann, D. Search for C 60 Fullerene in Char Produced on a Norway Spruce by Lightning. Fuller. Sci. Technol. 1998, 6 (6), 1079–1086 DOI: 10.1080/10641229809350257. (13) Benn, T.; Herckes, P.; Westerhoff, P. Fullerenes in Environmental Samples: C60 in Atmospheric Particulate Matter. In Comprehensive Analytical Chemistry; Elsevier, 2012; Vol. 59, pp 291–303. (14) Lee, J.; Song, W.; Jang, S. S.; Fortner, J. D.; Alvarez, P. J. J.; Cooper, W. J.; Kim, J.-H. Stability of Water-Stable C 60 Clusters to OH Radical Oxidation and Hydrated Electron Reduction. Environ. Sci. Technol. 2010, 44 (10), 3786–3792 DOI: 10.1021/es903550e. (15) Wu, J.; Benoit, D.; Lee, S. S.; Li, W.; Fortner, J. D. Ground State Reactions of nC 60 with Free Chlorine in Water. Environ. Sci. Technol. 2016, 50 (2), 721–731 DOI: 10.1021/acs.est.5b04368. (16) Hirsch, A.; Brettreich, M. Fullerenes: Chemistry and Reactions; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, FRG, 2004. (17) Heymann, D.; Bachilo, S. M.; Weisman, R. B.; Cataldo, F.; Fokkens, R. H.; Nibbering, N. M. M.; Vis, R. D.; Chibante, L. P. F. C 60 O 3 , a Fullerene Ozonide: Synthesis and Dissociation to C 60 O and O 2. J. Am. Chem. Soc. 2000, 122 (46), 11473–11479 DOI: 10.1021/ja0025651. (18) Malhotra, R.; Kumar, S.; Satyam, A. Ozonolysis of [60]fullerene. J. Chem. Soc. Chem. Commun. 1994, No. 11, 1339 DOI: 10.1039/c39940001339. (19) Lee, J.; Cho, M.; Fortner, J. D.; Hughes, J. B.; Kim, J.-H. Transformation of Aggregated C 60 in the Aqueous Phase by UV Irradiation. Environ. Sci. Technol. 2009, 43 (13), 4878– 4883 DOI: 10.1021/es8035972. (20) Arbogast, J. W.; Darmanyan, A. P.; Foote, C. S.; Diederich, F. N.; Whetten, R. L.; Rubin, Y.; Alvarez, M. M.; Anz, S. J. Photophysical properties of sixty atom carbon molecule (C60). J. Phys. Chem. 1991, 95 (1), 11–12 DOI: 10.1021/j100154a006. (21) Krusic, P. J.; Wasserman, E.; Parkinson, B. A.; Malone, B.; Holler, E. R.; Keizer, P. N.; Morton, J. R.; Preston, K. F. Electron spin resonance study of the radical reactivity of C60. J. Am. Chem. Soc. 1991, 113 (16), 6274–6275 DOI: 10.1021/ja00016a056. (22) Yamakoshi, Y.; Umezawa, N.; Ryu, A.; Arakane, K.; Miyata, N.; Goda, Y.; Masumizu, T.; Nagano, T. Active Oxygen Species Generated from Photoexcited Fullerene (C 60 ) as Potential Medicines: O 2 - • versus 1 O 2. J. Am. Chem. Soc. 2003, 125 (42), 12803–12809 DOI: 10.1021/ja0355574. (23) Lee, J.; Yamakoshi, Y.; Hughes, J. B.; Kim, J.-H. Mechanism of C 60 Photoreactivity in Water: Fate of Triplet State and Radical Anion and Production of Reactive Oxygen Species. Environ. Sci. Technol. 2008, 42 (9), 3459–3464 DOI: 10.1021/es702905g. (24) Utsunomiya, S.; Jensen, K. A.; Keeler, G. J.; Ewing, R. C. Uraninite and Fullerene in Atmospheric Particulates. Environ. Sci. Technol. 2002, 36 (23), 4943–4947 DOI: 10.1021/es025872a. (25) Sanchís, J.; Berrojalbiz, N.; Caballero, G.; Dachs, J.; Farré, M.; Barceló, D. Occurrence of Aerosol-Bound Fullerenes in the Mediterranean Sea Atmosphere. Environ. Sci. Technol. 2012, 46 (3), 1335–1343 DOI: 10.1021/es200758m. (26) Tiwari, A. J.; Morris, J. R.; Vejerano, E. P.; Hochella, M. F.; Marr, L. C. Oxidation of C 60 Aerosols by Atmospherically Relevant Levels of O 3. Environ. Sci. Technol. 2014, 48 (5), 2706–2714 DOI: 10.1021/es4045693.

22 ACS Paragon Plus Environment

Page 22 of 26

Page 23 of 26 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

443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487

ACS Earth and Space Chemistry

(27) Liu, Y.; Liggio, J.; Li, S.-M.; Breznan, D.; Vincent, R.; Thomson, E. M.; Kumarathasan, P.; Das, D.; Abbatt, J.; Antiñolo, M.; et al. Chemical and Toxicological Evolution of Carbon Nanotubes During Atmospherically Relevant Aging Processes. Environ. Sci. Technol. 2015, 150209062411004 DOI: 10.1021/es505298d. (28) Farré, M.; Barceló, D.; Farré, M. Analysis and risk of nanomaterials in environmental and food samples, 1. ed.; Comprehensive analytical chemistry; Elsevier: Amsterdam [u.a], 2012. (29) Katrib, Y.; Biskos, G.; Buseck, P. R.; Davidovits, P.; Jayne, J. T.; Mochida, M.; Wise, M. E.; Worsnop, D. R.; Martin, S. T. Ozonolysis of Mixed Oleic-Acid/Stearic-Acid Particles: Reaction Kinetics and Chemical Morphology. J. Phys. Chem. A 2005, 109 (48), 10910– 10919 DOI: 10.1021/jp054714d. (30) Smith, J. D.; Kroll, J. H.; Cappa, C. D.; Che, D. L.; Liu, C. L.; Ahmed, M.; Leone, S. R.; Worsnop, D. R.; Wilson, K. R. The heterogeneous reaction of hydroxyl radicals with submicron squalane particles: a model system for understanding the oxidative aging of ambient aerosols. Atmospheric Chem. Phys. 2009, 9 (9), 3209–3222 DOI: 10.5194/acp-93209-2009. (31) McNeill, V. F.; Yatavelli, R. L. N.; Thornton, J. A.; Stipe, C. B.; Landgrebe, O. Heterogeneous OH oxidation of palmitic acid in single component and internally mixed aerosol particles: vaporization and the role of particle phase. Atmospheric Chem. Phys. 2008, 8 (17), 5465–5476 DOI: 10.5194/acp-8-5465-2008. (32) Kessler, S. H.; Smith, J. D.; Che, D. L.; Worsnop, D. R.; Wilson, K. R.; Kroll, J. H. Chemical Sinks of Organic Aerosol: Kinetics and Products of the Heterogeneous Oxidation of Erythritol and Levoglucosan. Environ. Sci. Technol. 2010, 44 (18), 7005– 7010 DOI: 10.1021/es101465m. (33) Kwamena, N.-O. A.; Staikova, M. G.; Donaldson, D. J.; George, I. J.; Abbatt, J. P. D. Role of the Aerosol Substrate in the Heterogeneous Ozonation Reactions of SurfaceBound PAHs. J. Phys. Chem. A 2007, 111 (43), 11050–11058 DOI: 10.1021/jp075300i. (34) Kwamena, N.-O. A.; Thornton, J. A.; Abbatt, J. P. D. Kinetics of Surface-Bound Benzo[ a ]pyrene and Ozone on Solid Organic and Salt Aerosols. J. Phys. Chem. A 2004, 108 (52), 11626–11634 DOI: 10.1021/jp046161x. (35) Bedjanian, Y.; Nguyen, M. L. Kinetics of the reactions of soot surface-bound polycyclic aromatic hydrocarbons with O3. Chemosphere 2010, 79 (4), 387–393 DOI: 10.1016/j.chemosphere.2010.02.009. (36) Kang, E.; Root, M. J.; Toohey, D. W.; Brune, W. H. Introducing the concept of Potential Aerosol Mass (PAM). Atmospheric Chem. Phys. 2007, 7 (22), 5727–5744 DOI: 10.5194/acp-7-5727-2007. (37) Deguchi, S.; Alargova, R. G.; Tsujii, K. Stable Dispersions of Fullerenes, C 60 and C 70 , in Water. Preparation and Characterization. Langmuir 2001, 17 (19), 6013–6017 DOI: 10.1021/la010651o. (38) Kang, E.; Toohey, D. W.; Brune, W. H. Dependence of SOA oxidation on organic aerosol mass concentration and OH exposure: experimental PAM chamber studies. Atmospheric Chem. Phys. 2011, 11 (4), 1837–1852 DOI: 10.5194/acp-11-1837-2011. (39) Lambe, A. T.; Onasch, T. B.; Massoli, P.; Croasdale, D. R.; Wright, J. P.; Ahern, A. T.; Williams, L. R.; Worsnop, D. R.; Brune, W. H.; Davidovits, P. Laboratory studies of the chemical composition and cloud condensation nuclei (CCN) activity of secondary organic

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

488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531

(40)

(41)

(42)

(43)

(44)

(45)

(46) (47)

(48)

(49) (50)

(51)

aerosol (SOA) and oxidized primary organic aerosol (OPOA). Atmospheric Chem. Phys. Discuss. 2011, 11 (5), 13617–13653 DOI: 10.5194/acpd-11-13617-2011. Ortega, A. M.; Day, D. A.; Cubison, M. J.; Brune, W. H.; Bon, D.; de Gouw, J. A.; Jimenez, J. L. Secondary organic aerosol formation and primary organic aerosol oxidation from biomass-burning smoke in a flow reactor during FLAME-3. Atmospheric Chem. Phys. 2013, 13 (22), 11551–11571 DOI: 10.5194/acp-13-11551-2013. Li, R.; Palm, B. B.; Ortega, A. M.; Hlywiak, J.; Hu, W.; Peng, Z.; Day, D. A.; Knote, C.; Brune, W. H.; de Gouw, J. A.; et al. Modeling the Radical Chemistry in an Oxidation Flow Reactor: Radical Formation and Recycling, Sensitivities, and the OH Exposure Estimation Equation. J. Phys. Chem. A 2015, 119 (19), 4418–4432 DOI: 10.1021/jp509534k. Peng, Z.; Day, D. A.; Stark, H.; Li, R.; Lee-Taylor, J.; Palm, B. B.; Brune, W. H.; Jimenez, J. L. HOx radical chemistry in oxidation flow reactors with low-pressure mercury lamps systematically examined by modeling. Atmospheric Meas. Tech. 2015, 8 (11), 4863–4890 DOI: 10.5194/amt-8-4863-2015. Lambe, A. T.; Ahern, A. T.; Williams, L. R.; Slowik, J. G.; Wong, J. P. S.; Abbatt, J. P. D.; Brune, W. H.; Ng, N. L.; Wright, J. P.; Croasdale, D. R.; et al. Characterization of aerosol photooxidation flow reactors: heterogeneous oxidation, secondary organic aerosol formation and cloud condensation nuclei activity measurements. Atmospheric Meas. Tech. 2011, 4 (3), 445–461 DOI: 10.5194/amt-4-445-2011. Renbaum, L. H.; Smith, G. D. Artifacts in measuring aerosol uptake kinetics: the roles of time, concentration and adsorption. Atmospheric Chem. Phys. 2011, 11 (14), 6881–6893 DOI: 10.5194/acp-11-6881-2011. Mao, J.; Ren, X.; Brune, W. H.; Olson, J. R.; Crawford, J. H.; Fried, A.; Huey, L. G.; Cohen, R. C.; Heikes, B.; Singh, H. B.; et al. Airborne measurement of OH reactivity during INTEX-B. Atmospheric Chem. Phys. 2009, 9 (1), 163–173 DOI: 10.5194/acp-9163-2009. Seinfeld, J. H. Atmospheric chemistry and physics: from air pollution to climate change, 2nd ed.; J. Wiley: Hoboken, N.J, 2006. Wang, L.; Wu, R.; Xu, C. Atmospheric Oxidation Mechanism of Benzene. Fates of Alkoxy Radical Intermediates and Revised Mechanism. J. Phys. Chem. A 2013, 117 (51), 14163–14168 DOI: 10.1021/jp4101762. Zhang, Z.; Lin, L.; Wang, L. Atmospheric oxidation mechanism of naphthalene initiated by OH radical. A theoretical study. Phys. Chem. Chem. Phys. 2012, 14 (8), 2645 DOI: 10.1039/c2cp23271e. Wu, R.; Pan, S.; Li, Y.; Wang, L. Atmospheric Oxidation Mechanism of Toluene. J. Phys. Chem. A 2014, 118 (25), 4533–4547 DOI: 10.1021/jp500077f. Brant, J. A.; Labille, J.; Robichaud, C. O.; Wiesner, M. Fullerol cluster formation in aqueous solutions: Implications for environmental release. J. Colloid Interface Sci. 2007, 314 (1), 281–288 DOI: 10.1016/j.jcis.2007.05.020. Wu, J.; Alemany, L. B.; Li, W.; Petrie, L.; Welker, C.; Fortner, J. D. Reduction of Hydroxylated Fullerene (Fullerol) in Water by Zinc: Reaction and Hemiketal Product Characterization. Environ. Sci. Technol. 2014, 48 (13), 7384–7392 DOI: 10.1021/es5012912.

24 ACS Paragon Plus Environment

Page 24 of 26

Page 25 of 26 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

532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576

ACS Earth and Space Chemistry

(52) Wu, J.; Goodwin, D. G.; Peter, K.; Benoit, D.; Li, W.; Fairbrother, D. H.; Fortner, J. D. Photo-Oxidation of Hydrogenated Fullerene (Fullerane) in Water. Environ. Sci. Technol. Lett. 2014, 1 (12), 490–494 DOI: 10.1021/ez5003055. (53) Scrivens, W. A.; Tour, J. M.; Creek, K. E.; Pirisi, L. Synthesis of 14C-Labeled C60, Its Suspension in Water, and Its Uptake by Human Keratinocytes. J. Am. Chem. Soc. 1994, 116 (10), 4517–4518 DOI: 10.1021/ja00089a067. (54) Wei, X.; Wu, M.; Qi, L.; Xu, Z. Selective solution-phase generation and oxidation reaction of C60n− (n = 1,2) and formation of an aqueous colloidal solution of C60. J. Chem. Soc. Perkin Trans. 2 1997, No. 7, 1389–1394 DOI: 10.1039/a607336k. (55) Alargova, R. G.; Deguchi, S.; Tsujii, K. Stable Colloidal Dispersions of Fullerenes in Polar Organic Solvents. J. Am. Chem. Soc. 2001, 123 (43), 10460–10467 DOI: 10.1021/ja010202a. (56) Werner, H.; Wesemann, M.; Schlögl, R. Electronic Structure of C 60 (I 2 ) 1.88 : a Photoemission Study. Europhys. Lett. EPL 1992, 20 (2), 107–110 DOI: 10.1209/02955075/20/2/003. (57) Hou, W.-C.; Kong, L.; Wepasnick, K. A.; Zepp, R. G.; Fairbrother, D. H.; Jafvert, C. T. Photochemistry of Aqueous C 60 Clusters: Wavelength Dependency and Product Characterization. Environ. Sci. Technol. 2010, 44 (21), 8121–8127 DOI: 10.1021/es101230q. (58) Orrling, D.; Fitzgerald, E.; Ivanov, A.; Molina, M. Enhanced sulfate formation on ozoneexposed soot. J. Aerosol Sci. 2011, 42 (9), 615–620 DOI: 10.1016/j.jaerosci.2011.04.004. (59) Krusic, P. J.; Wasserman, E.; Keizer, P. N.; Morton, J. R.; Preston, K. F. Radical Reactions of C60. Science 1991, 254 (5035), 1183–1185 DOI: 10.1126/science.254.5035.1183. (60) Shiraiwa, M.; Ammann, M.; Koop, T.; Poschl, U. Gas uptake and chemical aging of semisolid organic aerosol particles. Proc. Natl. Acad. Sci. 2011, 108 (27), 11003–11008 DOI: 10.1073/pnas.1103045108. (61) McNeill, V. F.; Loerting, T.; Geiger, F. M.; Trout, B. L.; Molina, M. J. Hydrogen chloride-induced surface disordering on ice. Proc. Natl. Acad. Sci. 2006, 103 (25), 9422– 9427 DOI: 10.1073/pnas.0603494103. (62) Huthwelker, T.; Ammann, M.; Peter, T. The Uptake of Acidic Gases on Ice. Chem. Rev. 2006, 106 (4), 1375–1444 DOI: 10.1021/cr020506v. (63) Flückiger, B.; Chaix, L.; Rossi, M. J. Properties of the HCl/Ice, HBr/Ice, and H 2 O/Ice Interface at Stratospheric Temperatures (200 K) and Its Importance for Atmospheric Heterogeneous Reactions. J. Phys. Chem. A 2003, 107 (15), 2768–2768 DOI: 10.1021/jp0300400. (64) Klajn, R.; Bishop, K. J. M.; Grzybowski, B. A. Light-controlled self-assembly of reversible and irreversible nanoparticle suprastructures. Proc. Natl. Acad. Sci. 2007, 104 (25), 10305–10309 DOI: 10.1073/pnas.0611371104. (65) David, W. I. F.; Ibberson, R. M.; Dennis, T. J. S.; Hare, J. P.; Prassides, K. Structural Phase Transitions in the Fullerene C60. Europhys. Lett. EPL 1992, 18 (8), 735–736. (66) Pillar, E. A.; Camm, R. C.; Guzman, M. I. Catechol Oxidation by Ozone and Hydroxyl Radicals at the Air–Water Interface. Environ. Sci. Technol. 2014, 48 (24), 14352–14360 DOI: 10.1021/es504094x.

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

ACS Paragon Plus Environment

Page 26 of 26