Gas Phase Oxidation of Nicotine by OH Radicals: Kinetics

Jul 28, 2016 - Cigarette smoke is recognized as having harmful health effects for the smoker and for people breathing second-hand smoke. In an atmosph...
0 downloads 10 Views 739KB Size
Subscriber access provided by Northern Illinois University

Letter

Gas Phase Oxidation of Nicotine by OH Radicals: Kinetics, Mechanisms and Formation of HNCO Nadine Borduas, Jennifer Grace Murphy, Chen Wang, Gabriel da Silva, and Jonathan P.D. Abbatt Environ. Sci. Technol. Lett., Just Accepted Manuscript • DOI: 10.1021/acs.estlett.6b00231 • Publication Date (Web): 28 Jul 2016 Downloaded from http://pubs.acs.org on August 2, 2016

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

Environmental Science & Technology Letters 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 18

Environmental Science & Technology Letters

1

Gas Phase Oxidation of Nicotine by OH Radicals:

2

Kinetics, Mechanisms and Formation of HNCO

3

Nadine Borduas1*, Jennifer G. Murphy1, Chen Wang2, Gabriel da Silva3, Jonathan P. D. Abbatt1

4

1

5

Canada

6

2

7

of Toronto Scarborough, 1265 Military Trail, Toronto, ON, M1C 1A4, Canada

8

3

9

ABSTRACT

Department of Chemistry, University of Toronto, 80 St. George Street, Toronto, ON, M5S 3H6,

Department of Physical and Environmental Sciences and Department of Chemistry, University

Chemical and Biomolecular Engineering, University of Melbourne, Victoria 3010, Australia

10

Cigarette smoke is recognized as having harmful health effects for the smoker and for people

11

breathing second-hand smoke. In an atmospheric chemistry context however, little is known about

12

the fate of organic nitrogen compounds present in cigarette smoke. Indeed, the atmospheric

13

oxidation of nicotine, a major nitrogen-containing component of cigarette smoke, by OH radicals

14

has yet to be investigated. We report the first rate coefficient measurement between OH and

15

nicotine to be (8.38 ± 0.28) × 10-11 cm3 molec-1 s-1 at 298 ± 3 K. We use an online proton-transfer-

16

reaction mass spectrometer (PTR-MS) to quantify nicotine’s oxidation products, including

17

formamide and isocyanic acid (HNCO). We present the first evidence that HNCO is formed from

ACS Paragon Plus Environment

1

Environmental Science & Technology Letters

Page 2 of 18

18

nicotine’s gas phase oxidation and we highlight the potential for this toxic molecule to be an indoor

19

air pollutant after smoking has ended. Mechanistic pathways for the oxidation of nicotine by OH

20

radicals were investigated by theoretical calculations at the M06-2X level of theory and we find

21

that there are many competitive H-abstraction sites on nicotine. Our findings suggest that the

22

atmospheric removal of nicotine by OH radicals may compete with surface deposition and air

23

exchange, and may be a source of HNCO in indoor air.

24

TOC

25 26 27

INTRODUCTION

28

Smoking has well-documented negative health consequences for populations worldwide,

29

including cancer, pulmonary diseases and cardiovascular diseases.1 Exposure pathways of concern

30

include mainstream smoke, inhaled by the smoker, sidestream smoke (or second-hand smoke),

31

inhaled in the presence of a smoker, and third-hand smoke, inhaled near surfaces previously

32

exposed to cigarette smoke.2 Both mainstream and sidestream smoking increase cardiovascular

33

disease risk by promoting atherosclerosis and blood clot formation.1 Nicotine is the major organic

34

nitrogen compound emitted from burning tobacco cigarettes. It can be present at levels up to 3 mg

35

per cigarette depending on the cigarette brand, and thus may account for ~1.5% of a cigarette’s

ACS Paragon Plus Environment

2

Page 3 of 18

Environmental Science & Technology Letters

36

weight.3-5 Nicotine is a known health hazard and causes addiction, but is also used as a therapeutic

37

for Parkinson and Alzheimer’s diseases.5

38

The atmospheric fate of nicotine has been assumed to be governed by its deposition to

39

surfaces.6,7 Based on the current literature, gas-phase nicotine is expected to be rapidly removed

40

by sorption to indoor surfaces, and re-emitted during ventilation, leading to third hand smoke

41

exposure and to concentrations up to ~10% of those observed during the smoking phase indoors.8

42

Nicotine adsorption rate coefficients to surfaces in an indoor environment range from 1.5 – 6 hours-

43

1

44

heterogeneous oxidation by ozone has been previously studied at high ozone levels and found to

45

proceed on timescales of a week or so.9 Surprisingly, its rate coefficient with OH radicals has yet

46

to be reported. Nicotine’s gas phase oxidation may compete with deposition and air exchange rates

47

in indoor environments despite expected lower concentrations of OH radicals indoor (~5 × 105

48

molec cm-3) compared to outdoor air (~2 × 106 molec cm-3).10,11 In an outdoor context, nicotine

49

emissions may be increasing through regulations requiring smokers to smoke outside, in which

50

case nicotine may be oxidized in a higher OH radical concentration environment. In addition,

51

electronic cigarettes generate fewer particles, allowing nicotine to partition and/or remain in the

52

gas phase.12

and 95-99% is reported to be partitioned to the sorbed phase at equilibrium.6,7 Nicotine’s

53

Sidestream smoke and nicotine exposure are of concern, but the aging of both the smoke and

54

nicotine may also have negative health and environmental impacts. Nicotine has been found to

55

contribute significantly (with yields of 4-9%) to the formation of secondary organic aerosol (SOA)

56

through reaction with ozone.13 SOA formation has also been observed by sorbed nicotine from

57

cellulose, cotton and paper surfaces reacting with ozone in the presence of NOx and varying RH.14

58

In addition, residual nicotine from tobacco smoke sorbed to indoor surfaces may react with

ACS Paragon Plus Environment

3

Environmental Science & Technology Letters

Page 4 of 18

59

ambient nitrous acid (HONO) to form carcinogenic tobacco-specific nitrosamines.15 These recent

60

studies raise concerns about exposure to tobacco smoke residue, through second- and third-hand

61

smoke.

62

In an effort to contribute to the growing knowledge of cigarette smoke from an atmospheric

63

chemistry perspective, we provide the first measurement of nicotine’s rate coefficient with OH

64

radicals at room temperature. Also, for the first time, we quantify formamide and HNCO as

65

oxidation products of this alkaloid and further support nicotine’s oxidation mechanism with DFT

66

calculations. The observations of HNCO formation are especially important given this

67

compound’s recognized toxicity. HNCO is known to cause carbamylation which may

68

subsequently lead to illnesses like rheumatoid arthritis, cataracts and cardiovascular diseases.16-19

69

HNCO has also recently been measured in ambient air at concentrations up to 1 ppb, high enough

70

to be of health concern.20 The fast gas phase loss kinetics of nicotine and the formation of toxic

71

products contribute to our understanding of the fate of cigarette smoke and to the production of

72

secondary pollutants, especially in an indoor environment.

73 74 75

MATERIALS AND METHODS 1. Kinetic and product study of nicotine

76

The kinetics of nicotine OH oxidation were studied using a quadrupole proton-transfer-reaction

77

mass spectrometer (PTR-MS) (Ionicon Analytik GmbH, Innsbruck Austria)21 connected to a 1 m3

78

Teflon FEP film (American Durafilm) bag (Ingeniven) and mounted on a Teflon-coated frame,

79

with no fan. The chamber is also surrounded by 24 UVB lamps (Microlites Scientific) centered at

80

310 nm used to photolyse hydrogen peroxide as the OH radical source. The signals of the

ACS Paragon Plus Environment

4

Page 5 of 18

Environmental Science & Technology Letters

81

compounds of interest were normalized to the H318O+ ion, m/z 21. Experiments were conducted

82

under dry (< 1 %) and NOx-free conditions (see SI for further details).

83

For the calibration of the PTR-MS, known volumes of a 0.7 M aqueous solution of nicotine

84

(Sigma-Aldrich, > 99% purity) were injected into the chamber through a glass tube connection. A

85

flow of pure air from an AADCO 737-series generator passed through the glass tube adding

86

nicotine to the chamber in mixing ratios up to ~700 ppbv. Nicotine was monitored at its protonated

87

molecular ion at m/z 163, and had a linear calibration (Figure S1), a sensitivity of 2 ncps/ppbv

88

(calculated using Equation S1)21 and a LOD of 2.0 ppbv (3). Furthermore, the m/z 44 and m/z 46

89

PTR-MS signals were assigned to HNCO and formamide. Calibration details of the products are

90

detailed in the SI.

91

A typical oxidation experiment used 1,3,5-trimethylbenzene (TMB) (Sigma-Aldrich, 98%) as

92

the reference compound for the relative rate kinetics, detected by the PTR-MS at m/z 121.22 An

93

aqueous solution of 30 % hydrogen peroxide bubbled into the chamber generated OH radicals

94

upon UVB light exposure. The hydrogen peroxide was continuously injected inside the chamber

95

to ensure a steady state concentration of OH radicals and thus pseudo-first order kinetics (see

96

Figure S3). The OH radical concentration was ~ 107 molec cm-3, estimated by the decay of the

97

reference compound. Refer to the SI for the details on control experiments conducted.

98

2. Theoretical method

99

To investigate the mechanism of the gas-phase reaction of nicotine with OH radicals,

100

computational DFT methods were employed using the Gaussian 09 code.23 Structures of the

101

reactants and transition states were optimized using the M06-2X density functional method with

102

the 6-311G+(d,p) basis set.24 The energies reported are 0 K enthalpies and are calculated from the

ACS Paragon Plus Environment

5

Environmental Science & Technology Letters

Page 6 of 18

103

sum of the electronic energy and the zero point energy. Optimized geometries are provided in the

104

SI.

105

The transition state energy represents the energy difference between the transition state and the

106

sum of the reactants’ energies. The assumption is that the energy of the pre-complex formation of

107

nicotine with OH radicals does not affect the rate of the reaction, i.e. we report transition state

108

energy relative to the energy of the reactants as an indication of whether the mechanism is likely

109

to occur readily.

110 111 112

RESULTS AND DISCUSSION 1. Kinetics results

113

The rate coefficient for nicotine’s gas phase reaction with OH radicals was measured using the

114

relative rate kinetics method.25 Nicotine was detected in the chamber, but required the use of

115

hundreds of ppbv for quantifiable decays. Compared to our previous experience with gas phase

116

amines, the chamber did not require prior conditioning for a reproducible nicotine signal to be

117

observed (see nicotine signal in Figure S3).22 Plotting the natural logarithm of the decay of nicotine

118

as a function of the natural logarithm of the decay of the reference compound as in Figure 1 yields

119

a linear regression in which the slope represents the ratio of the nicotine’s rate coefficient to TMB’s

120

rate coefficient. Figure 1 represents the data and fit from a single experiment, and the experiment

121

was performed in triplicate. Based on TMB’s previously reported room temperature rate

122

coefficient of (5.73 ± 0.53) × 10-11 cm3 molec-1 s-1,26 we convert each regression to its

123

corresponding rate coefficient for nicotine. The average of these triplicate rate coefficients leads

124

to a value of (8.38 ± 0.28) × 10-11 cm3 molec-1 s-1 for nicotine + OH radicals at 298 ± 3 K where

125

the standard deviation represents the variability between experiments. Since the precision between

ACS Paragon Plus Environment

6

Page 7 of 18

Environmental Science & Technology Letters

126

replicated experiments is good (3%), the uncertainty for nicotine’s rate coefficient is dominated

127

by the uncertainty in TMB’s rate coefficient, leading to an overall uncertainty closer to 10%. If we

128

assume an indoor environment with an OH radical concentration as high as 5 × 105 molec cm-3,10

129

nicotine would have a lifetime of 6.6 hours (or longer with lower averaged OH concentrations)27,

130

and a shorter lifetime outdoors where OH concentrations are expected to be higher during the

131

daytime.

132

The room temperature rate coefficient measured for nicotine can be compared to other amines’

133

reactivity towards OH radicals. Indeed, the majority of alkylamines have rate coefficients between

134

6-9 × 10-11 cm3 molec-1 s-1.28-31 The rate coefficient of heterogeneous nicotine with gas phase ozone

135

has been reported to be 0.035 ± 0.015 min-1 in the presence of 245 ppmv of ozone.9 This rate

136

coefficient is not directly transferable to the gas phase, but leads us to believe that the atmospheric

137

oxidation of nicotine will be dominated by OH radicals rather than by ozone.

138

Concerns over whether the nicotine decay observed was occurring in the gas phase versus on

139

the wall/particle surface are justified. However, we argue that the majority of the decay was

140

occurring in the gas phase during the first hour of oxidation. First, the decay rate of nicotine in a

141

flushing control experiment matched the purging rate of the chamber, suggesting minimal

142

desorption from the wall (Figure S2). Second, particle formation is inevitable during oxidation,

143

but the exponential decay of nicotine and TMB in Figure S3 and their linear relationship in Figure

144

1 indicates that nicotine does not have an additional and/or growing sink as the experiment

145

progresses. Further discussion is included in the SI.

146

The fate of gas phase nicotine in a typical indoor environment depends on its deposition to

147

surfaces, the air exchange rate of the room/building and its chemical reactivity. Studies have

148

previously suggested that the majority of nicotine is removed via deposition, and may later

ACS Paragon Plus Environment

7

Environmental Science & Technology Letters

Page 8 of 18

149

revolatilize.7,8,14 Air exchange rates in residential homes range between 0.05 – 5 h-1.32 Our results

150

suggest that chemical reactivity with OH radicals may compete with ventilation for removal in

151

particularly well-sealed buildings with low air exchange rates. The gas phase reactivity of nicotine

152

will also apply to deposited molecules that repartition to the gas phase.

153 154

Figure 1: Plot of the natural logarithm of the decay of nicotine against the natural logarithm

155

of the decay of the reference compound TMB. Each point was taken at ~ 5 s intervals by the

156

PTR-MS and the linear regression is represented as the black line. The nicotine decay plotted

157

here was taken from 0.08 – 1.2 h from Figure S3A. (See SI for further details on this

158

regression.)

159 160

2. Product study

161

Nicotine reacts quickly in the presence of OH radicals and so identifying its oxidation products

162

is important to better assess its health and environmental impact in the context of cigarette smoke.

163

Figure 2 depicts the evolution of nicotine and its nitrogenated products as a function of time,

164

measured by the PTR-MS. The major oxidation products containing nitrogen that could be

165

confidently assigned were formamide and HNCO. If we assume that all the nicotine injected was

ACS Paragon Plus Environment

8

Page 9 of 18

Environmental Science & Technology Letters

166

oxidized, we can report lower limit yields of 5 % for both formamide and HNCO after 2 h of

167

photo-oxidation. The yield of formamide is similar to measured yields from other amines like

168

monoethanolamine (MEA).22 We note that these yields are likely lower limits since the products

169

do appear to continue growing after the disappearance of nicotine (Figure S4). We also add that

170

these yields are not directly transferable to the real atmosphere as we operate in a NOx-free

171

environment. The production of HNCO from nicotine’s photo-oxidation was also confirmed by

172

appending an acetate reagent ion chemical ionization mass spectrometer to the chamber, an

173

instrument described by Borduas et al.,33 which does not have interferences at HNCO’s molecular

174

ion. Despite uncertainties in HNCO quantification (detailed in the SI), the observation of HNCO

175

production from nicotine photo-oxidation is important, and may contribute to the negative health

176

effects observed with smoking tobacco cigarettes.

177

By plotting the mixing ratios of formamide and HNCO over time as a function of the decaying

178

nicotine mixing ratio (Figure S5), it is clear that these compounds are not first order generation

179

products. In fact, the exponential curve in Figure S5 suggests they are at least second or third

180

generation products, implying a complex mechanism of formation. Recent studies have shown that

181

HNCO is the major product of formamide photo-oxidation,34-36 yet HNCO must have other sources

182

based on the early onset of its production during the nicotine’s oxidation (see Figures 2 and S4).

183

A similar conclusion was also made by Link et al. in their study of photo-oxidation of diesel

184

exhaust.37 Furthermore, Roberts et al. detected high levels of HNCO from cigarette smoke in a

185

preliminary study.20 Our study extends Roberts et al.’s observation; even after smoking has ended,

186

there are secondary sources of HNCO, at least from nicotine, with important implications for

187

indoor air quality. HNCO’s dominant ambient atmospheric fate is to partition to the aqueous phase,

ACS Paragon Plus Environment

9

Environmental Science & Technology Letters

Page 10 of 18

188

where it may have lifetimes from hours to months depending on pH,33 but yet unknown lifetimes

189

in the indoor environment.

190

Finally, the signal at m/z 42 was assigned to acetonitrile and detected with a yield of ~ 1 %.

191

Acetonitrile is typically known as a biomass tracer in ambient air and is expected to be formed

192

from the oxidation of organic nitrogen.38 700

70 nicotine isocyanic acid formamide acetonitrile

500

60

50

400

40

300

30

200

20

100

10

0

ppbv of products

ppbv of nicotine

600

0 0.0

193

0.4

0.8 1.2 hours of reaction

1.6

2.0

194

Figure 2: The evolution of nicotine and its oxidation products upon exposure to OH radicals.

195

The m/z with their assignments are m/z 163 (nicotine), m/z 44 (isocyanic acid), m/z 46

196

(formamide), and m/z 42 (acetonitrile). See also Figure S4.

197 198

3. Mechanism of oxidation

199

To explore the mechanism of oxidation of nicotine, as well as to gain insight into the first-

200

generation oxidation products, we employ computational chemistry to calculate the transition state

201

energies of nicotine + OH radicals relative to the energies of the reactants. Nicotine’s benzylic

202

C−H bond, the most electron rich bond of the molecule, is likely to be the most reactive towards

203

OH (Figure 3 and Figure S6). Its abstraction has been observed as the primary pathway in

ACS Paragon Plus Environment

10

Page 11 of 18

Environmental Science & Technology Letters

204

nicotine’s aqueous phase oxidation.39 We calculate a nicotine-OH radical pre-complex at – 3.6

205

kcal mol-1 below the entrance energies of the reactants with a transition state only 0.4 kcal mol -1

206

above the energy of the pre-complex (Figure 3). These energies translate to a facile C−H

207

abstraction and are consistent with the large rate coefficient we measured.

208

Nonetheless, there exist fourteen C−H bonds (nine of them unique) in nicotine which could react

209

with OH radicals, in addition to potential OH-addition sites on the pyridine ring. To identify which

210

C−H abstraction governs nicotine’s reactivity, we iteratively calculated each C−H bond’s

211

transition state energy. These results are presented in Figure S6 and in the inset of Figure 3 and

212

suggest that many C−H abstraction mechanisms are competitive for reaction with OH radicals.

213

Thus, the oxidation mechanisms of nicotine appear to be complex and occur at several C−H

214

abstraction sites (in blue in the inset of Figure 3). Furthermore, we highlight unique reactivity on

215

nicotine’s pyridine ring where the OH addition mechanism is deactivated (see SI and Figure S7).40

216

It becomes difficult to assign a single mechanistic pathway for the production of formamide and

217

HNCO. We estimate that at least two closed-shell intermediates are likely prior to formamide

218

production based on calculations of subsequent reactions of the benzylic radical product (not

219

shown). We also cannot differentiate if the nitrogen in formamide originates from the N−CH3

220

group or the pyridine ring N. Further studies with isotopically labelled nitrogen might help in

221

elucidating and differentiating mechanisms leading to formamide and HNCO.

ACS Paragon Plus Environment

11

Environmental Science & Technology Letters

Page 12 of 18

222 223

Figure 3: Theoretical energy diagram for nicotine + OH. Energies are 0 K enthalpies in kcal

224

mol-1, at the M06-2X level of theory and 6-311G+(d,p) basis set. Inset: lowest transition state

225

energies for the C−H bond abstraction and OH addition mechanism. Likely reaction sites

226

are colored in blue and unlikely reaction mechanisms are colored in red. Also see Figure S3.

227 228

AUTHOR INFORMATION

229

Corresponding Author

230

*[email protected], + 1 416 946 0260, @nadineborduas

231

Author Contributions

ACS Paragon Plus Environment

12

Page 13 of 18

Environmental Science & Technology Letters

232

The manuscript was written through contributions of all authors. All authors have given approval

233

to the final version of the manuscript.

234

Funding Sources

235

NB is grateful to the University of Toronto for the Adel Sedra Graduate Award as well as for the

236

Special Opportunity Travel Grant. The authors also acknowledge funding from CFI, NSERC and

237

the Sloan Foundation. GdS is supported by the Australian Research Council Future Fellowship

238

(FT130101304) and Discovery Projects (DP110103889) programs.

239

ACKNOWLEDGMENT

240

We thank Environment Canada for lending us their VOC standard canister and John Wang for

241

assistance using the VOC canister.

242

ASSOCIATED CONTENT

243

Supporting Information Available: Detailed information on control experiments, theoretical results

244

and optimized geometries.

245

REFERENCES

246 247 248 249

References 1. Eriksen, M.; Mackay, J.; Schluger, N.; Gomeshtapeh, F. I.; Drope, J., Eds.; In The Tobacco Atlas; American Cancer Society: Atlanta, Georgia, 2015; pp 8-14. 2. Guerin, M. R.; Higgins, C. E.; Jenkins, R. A. Measuring environmental emissions from

250

tobacco combustion: Sidestream cigarette smoke literature review. Atmos. Environ. 1987,

251

21, 291-297.

252 253

3. Charles, S. M.; Batterman, S. A.; Jia, C. Composition and emissions of VOCs in main- and side-stream smoke of research cigarettes. Atmos. Environ. 2007, 41, 5371-5384.

ACS Paragon Plus Environment

13

Environmental Science & Technology Letters

254

Page 14 of 18

4. Charles, S. M.; Jia, C.; Batterman, S. A.; Godwin, C. VOC and particulate emissions from

255

commercial cigarettes: Analysis of 2,5-DMF as an ETS tracer. Environ. Sci. Technol. 2008,

256

42, 1324-1331.

257 258 259

5. Hukkanen, J.; Jacob, P.; Benowitz, N. L. Metabolism and disposition kinetics of nicotine. Pharmacol. Rev. 2005, 57, 79-115. 6. Singer, B. C.; Hodgson, A. T.; Nazaroff, W. W. Gas-phase organics in environmental tobacco

260

smoke: 2. Exposure-relevant emission factors and indirect exposures from habitual smoking.

261

Atmos. Environ. 2003, 37, 5551-5561.

262 263

7. Singer, B. C.; Revzan, K. L.; Hotchi, T.; Hodgson, A. T.; Brown, N. J. Sorption of organic gases in a furnished room. Atmos. Environ. 2004, 38, 2483-2494.

264

8. Petrick, L. M.; Sleiman, M.; Dubowski, Y.; Gundel, L. A.; Destaillats, H. Tobacco smoke

265

aging in the presence of ozone: A room-sized chamber study. Atmos. Environ. 2011, 45,

266

4959-4965.

267 268 269

9. Petrick, L.; Destaillats, H.; Zouev, I.; Sabach, S.; Dubowski, Y. Sorption, desorption, and surface oxidative fate of nicotine. Phys. Chem. Chem. Phys. 2010, 12, 10356-10364. 10. Gómez Alvarez, E.; Amedro, D.; Afif, C.; Gligorovski, S.; Schoemaecker, C.; Fittschen, C.;

270

Doussin, J.; Wortham, H. Unexpectedly high indoor hydroxyl radical concentrations

271

associated with nitrous acid. Proc. Natl. Acad. Sci. U. S. A. 2013, 110, 13294-13299.

272

11. Jacobs, D. J. Oxidizing Power of the Troposphere. In Introduction to Atmospheric

273

ChemistryPrinceton University Press: Princeton, New Jersey, 1999; pp 200-219.

274

12. Bush, D.; Goniewicz, M. L. A pilot study on nicotine residues in houses of electronic

275

cigarette users, tobacco smokers, and non-users of nicotine-containing products. Int. J. Drug

276

Policy 2015, 26, 609-611.

277

13. Sleiman, M.; Destaillats, H.; Smith, J. D.; Liu, C.; Ahmed, M.; Wilson, K. R.; Gundel, L. A.

278

Secondary organic aerosol formation from ozone-initiated reactions with nicotine and

279

secondhand tobacco smoke. Atmos. Environ. 2010, 44, 4191-4198.

ACS Paragon Plus Environment

14

Page 15 of 18

Environmental Science & Technology Letters

280

14. Petrick, L. M.; Svidovsky, A.; Dubowski, Y. Thirdhand smoke: Heterogeneous oxidation of

281

nicotine and secondary aerosol formation in the indoor environment. Environ. Sci. Technol.

282

2011, 45, 328-333.

283

15. Sleiman, M.; Gundel, L. A.; Pankow, J. F.; Jacob, P.; Singer, B. C.; Destaillats, H. Formation

284

of carcinogens indoors by surface-mediated reactions of nicotine with nitrous acid, leading

285

to potential thirdhand smoke hazards. Proc. Natl. Acad. Sci. U. S. A. 2010, 107, 6576-6581.

286

16. Wang, Z.; Nicholls, S. J.; Rodriguez, E. R.; Kummu, O.; Horkko, S.; Barnard, J.; Reynolds,

287

W. F.; Topol, E. J.; DiDonato, J. A.; Hazen, S. L. Protein carbamylation links inflammation,

288

smoking, uremia and atherogenesis. Nat. Med. 2007, 13, 1176-1184.

289

17. Mydel, P.; Wang, Z.; Brisslert, M.; Hellvard, A.; Dahlberg, L. E.; Hazen, S. L.; Bokarewa,

290

M. Carbamylation-dependent activation of T cells: A novel mechanism in the pathogenesis

291

of autoimmune arthritis. J. Immunol. 2010, 184, 6882-6890.

292

18. Jaisson, S.; Pietrement, C.; Gillery, P. Carbamylation-derived products: Bioactive

293

compounds and potential biomarkers in chronic renal failure and atherosclerosis. Clin.

294

Chem. 2011, 57, 1499-1505.

295 296

19. Verbrugge, F. H.; Tang, W. H. W.; Hazen, S. L. Protein carbamylation and cardiovascular disease. Kidney Int. 2015, 88, 474-478.

297

20. Roberts, J. M.; Veres, P. R.; Cochran, A. K.; Warneke, C.; Burling, I. R.; Yokelson, R. J.;

298

Lerner, B.; Gilman, J. B.; Kuster, W. C.; Fall, R.; de Gouw, J. Isocyanic acid in the

299

atmosphere and its possible link to smoke-related health effects. Proc. Natl. Acad. Sci. U. S.

300

A. 2011, 108, 8966-8971.

301

21. De Gouw, J.; Warneke, C. Measurements of volatile organic compounds in the Earth's

302

atmosphere using proton-transfer-reaction mass spectrometry. Mass Spectrom. Rev. 2007,

303

26, 223-257.

304

22. Borduas, N.; Abbatt, J. P. D.; Murphy, J. G. Gas phase oxidation of monoethanolamine

305

(MEA) with OH radical and ozone: kinetics, products, and particles. Environ. Sci. Technol.

306

2013, 47, 6377-6383.

ACS Paragon Plus Environment

15

Environmental Science & Technology Letters

307

Page 16 of 18

23. Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J.

308

R.; Montgomery, J. J. A.; Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S.

309

S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G.

310

A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.;

311

Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H.

312

P.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.;

313

Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.;

314

Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich,

315

S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari,

316

K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.;

317

Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.;

318

Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M.

319

W.; Johnson, B.; Chen, W.; Wong, M. W.; Gonzalez, C.; Pople, J. A. Gaussian 09, Revision

320

B.01. 2010.

321

24. Zhao, Y.; Truhlar, D. The M06 suite of density functionals for main group thermochemistry,

322

thermochemical kinetics, noncovalent interactions, excited states, and transition elements:

323

Two new functionals and systematic testing of four M06-class functionals and 12 other

324

functionals. Theor. Chem. Acc. 2008, 120, 215-241.

325

25. Atkinson, R. Kinetics and mechanisms of the gas-phase reactions of the hydroxyl radical

326

with organic compounds under atmospheric conditions. Chem. Rev. 1986, 86, 69-201.

327

26. Kramp, F.; Paulson, S. E. On the uncertainties in the rate coefficients for OH reactions with

328

hydrocarbons, and the rate coefficients of the 1,3,5-trimethylbenzene and m-xylene

329

reactions with OH radicals in the gas phase. J. Phys. Chem. A 1998, 102, 2685-2690.

330 331 332 333

27. Weschler, C. J.; Shields, H. C. Production of the hydroxyl radical in indoor air. Environ. Sci. Technol. 1996, 30, 3250-3258. 28. Ge, X.; Wexler, A. S.; Clegg, S. L. Atmospheric amines - Part I. A review. Atmos. Environ. 2011, 45, 524-546.

ACS Paragon Plus Environment

16

Page 17 of 18

334 335 336

Environmental Science & Technology Letters

29. Nielsen, C. J.; Herrmann, H.; Weller, C. Atmospheric chemistry and environmental impact of the use of amines in carbon capture and storage (CCS). Chem. Soc. Rev. 2012, 6684-6704. 30. Onel, L.; Thonger, L.; Blitz, M. A.; Seakins, P. W.; Bunkan, A. J. C.; Solimannejad, M.;

337

Nielsen, C. J. Gas-phase reactions of OH with methyl amines in the presence or absence of

338

molecular oxygen. An experimental and theoretical Study. J. Phys. Chem. A 2013, 117,

339

10736-10745.

340

31. Onel, L.; Dryden, M.; Blitz, M. A.; Seakins, P. W. Atmospheric oxidation of piperazine by

341

OH has a low potential to form carcinogenic compounds. Environ. Sci. Technol. Lett. 2014,

342

1, 367-371.

343

32. Breen, M. S.; Breen, M.; Williams, R. W.; Schultz, B. D. Predicting residential air exchange

344

rates from questionnaires and meteorology: Model evaluation in Central North Carolina.

345

Environ. Sci. Technol. 2010, 44, 9349-9356.

346

33. Borduas, N.; Place, B.; Wentworth, G. R.; Abbatt, J. P. D.; Murphy, J. G. Solubility and

347

reactivity of HNCO in water: insights into HNCO's fate in the atmosphere. Atmos. Chem.

348

Phys. 2016, 16, 703-714.

349 350 351

34. Barnes, I.; Solignac, G.; Mellouki, A.; Becker, K. H. Aspects of the atmospheric chemistry of amides. ChemPhysChem 2010, 11, 3844-3857. 35. Borduas, N.; da Silva, G.; Murphy, J. G.; Abbatt, J. P. D. Experimental and theoretical

352

understanding of the gas phase oxidation of atmospheric amides with OH radicals: kinetics,

353

products, and mechanisms. J. Phys. Chem. A 2015, 119, 4298-4308.

354

36. Bunkan, A. J. C.; Mikoviny, T.; Nielsen, C. J.; Wisthaler, A.; Zhu, L. Experimental and

355

theoretical study of the OH-initiated photo-oxidation of formamide. J. Phys. Chem. A 2016,

356

120, 1222-1230.

357

37. Link, M. F.; Friedman, B.; Fulgham, R.; Brophy, P.; Galang, A.; Jathar, S. H.; Veres, P.;

358

Roberts, J. M.; Farmer, D. K. Photochemical processing of diesel fuel emissions as a large

359

secondary source of isocyanic acid (HNCO). Geophys. Res. Lett. 2016, 43, 4033-4041.

ACS Paragon Plus Environment

17

Environmental Science & Technology Letters

360

38. Yuan, B.; Liu, Y.; Shao, M.; Lu, S.; Streets, D. G. Biomass burning contributions to ambient

361

VOCs species at a receptor site in the Pearl River Delta (PRD), China. Environ. Sci.

362

Technol. 2010, 44, 4577-4582.

363 364 365

Page 18 of 18

39. Kosno, K.; Janik, I.; Celuch, M.; Mirkowski, J.; Kisala, J.; Pogocki, D. The role of pH in the mechanism of OH radical induced oxidation of nicotine. Isr. J. Chem. 2014, 54, 302-315. 40. da Silva, G.; Bozzelli, J. W.; Asatryan, R. Hydroxyl radical initiated oxidation of s-triazine:

366

Hydrogen abstraction is faster than hydroxyl addition. J. Phys. Chem. A 2009, 113, 8596-

367

8606.

368

ACS Paragon Plus Environment

18