OH-Initiated Oxidation of Acetylacetone: Implications for Ozone and

Aug 30, 2018 - ... by Reactive Uptake of Methylglyoxal and Photolytic Cloud Cycling. ...... Software Manager; University of Minnesota: Minneapolis, MN...
0 downloads 0 Views 2MB Size
Subscriber access provided by Kaohsiung Medical University

Environmental Processes

OH-initiated oxidation of acetylacetone: Implications for ozone and secondary organic aerosol formation Yuemeng Ji, Jun Zheng, Dandan Qin, Yixin Li, Yanpeng Gao, Meijing Yao, Xingyu Chen, Guiying Li, Taicheng An, and Renyi Zhang Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b03972 • Publication Date (Web): 30 Aug 2018 Downloaded from http://pubs.acs.org on August 31, 2018

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

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

Page 1 of 32

Environmental Science & Technology

1

OH-initiated oxidation of acetylacetone: Implications for ozone and secondary organic

2

aerosol formation

3

Yuemeng Jia,1, Jun Zheng,b,1 Dandan Qin,a Yixin Li,c Yanpeng Gaoa, Meijing Yaoa, Xingyu

4

Chena, Guiying Lia, Taicheng Ana,*, and Renyi Zhangc,*

5

a

6

Environmental Science and Engineering, Institute of Environmental Health and Pollution

7

Control, Guangdong University of Technology, Guangzhou 510006, China b

8 9

Guangzhou Key Laboratory of Environmental Catalysis and Pollution Control, School of

Collaborative Innovation Center of Atmospheric Environment and Equipment Technology, Nanjing University of Information Science & Technology, Nanjing 210044, China c

10 11

Department of Atmospheric Sciences and Department of Chemistry, Texas A&M University, College Station, TX 77843

12

1

13

ABSTRACT. Acetylacetone (AcAc) is a common atmospheric oxygenated volatile organic

14

compound (OVOC) due to broad industrial applications, but its atmospheric oxidation

15

mechanism is not fully understood. We investigate the mechanism, kinetics, and atmospheric

16

fate of the OH-initiated oxidation for the enolic and ketonic isomers of AcAc using quantum

17

chemical and kinetic rate calculations. OH addition to enol-AcAc is more favorable than

18

addition to keto-AcAc, with the total rate constant of 1.69 × 10-13 exp(1935/T) cm3 molecule-1

19

s-1 over the temperature range of 200-310 K. For the reaction of the enol-AcAc with OH, the

20

activation energies of H-abstraction are at least 4 kcal mol-1 higher than those of OH-

21

addition, and the rate constants for OH-addition are by 2‒3 orders of magnitude higher than

Both authors contributed equally to this work

1

ACS Paragon Plus Environment

Environmental Science & Technology

22

those for H-abstraction. Oxidation of AcAc is predicted to yields significant amounts of

23

acetic acid and methylglyoxal, larger than those are currently recognized. A lifetime of less

24

than a few hours for AcAc is estimated throughout the tropospheric conditions. In addition,

25

we present field measurements in Beijing and Nanjing, China, showing significant

26

concentrations of AcAc in the two urban locations. Our results reveal that the OH-initiated

27

oxidation of AcAc contributes importantly to ozone and SOA formation under polluted

28

environments.

29

KEYWORDS. Oxygenated volatile organic compounds; atmospheric chemistry; mechanism

30

and kinetics; secondary organic aerosol; ozone

2

ACS Paragon Plus Environment

Page 2 of 32

Page 3 of 32

Environmental Science & Technology

31

INTRODUCTION

32

Ketones represent an important class of oxygenated volatile organic compounds (OVOCs)

33

and are emitted from natural and anthropogenic sources, e.g., from industrial activities and

34

oxidation reactions of both biogenic and anthropogenic hydrocarbons.1-5 The atmospheric

35

oxidation of ketones has been identified as important sources for HOx radical species (i.e.,

36

OH and HO2)6,7 and secondary organic aerosol (SOA),6,8 profoundly impacting air quality,

37

human health, and climate. Acetylacetone (also known as 2,4-pentanedione or AcAc) is

38

highly reactive and broadly used for various industrial applications. For example, AcAc is an

39

important reagent in preparation of chelate compounds for a wide range of transition

40

metals,9,10 an industrial additive,11 and a building block for synthesis of heterocyclic

41

compounds and raw materials for sulfonamide drugs.12 In developed countries such as Japan,

42

the United States, and Europe, the global capacity of AcAc is estimated to be approximately

43

20000 t a-1,13 mainly from industrial activities, while its formation from in situ atmospheric

44

photochemical production is negligible. Furthermore, emissions of AcAc in developing

45

countries (such as China) are anticipated to be substantially increased because of their rapid

46

industrialization and economic development.14,15

47

AcAc is a prototype of β-diketone that exists in the two tautomeric forms: the enol form

48

contains an intramolecular hydrogen bond and resonance stabilization through a conjugated

49

π-system, whereas the diketo form contains two carbonyl groups with an ∼140° dihedral

50

angle between the oxygens. The atmospheric oxidation mechanism of AcAc is complex,

3

ACS Paragon Plus Environment

Environmental Science & Technology

51

including multiple pathways and steps. Oxidation of AcAc is mainly initiated by the hydroxyl

52

radical (OH).13,16-18 An earlier experimental study by Zhou et al. has evaluated the

53

atmospheric chemistry for AcAc; using a relative kinetic method the authors determined the

54

temperature dependent rate coefficients over the temperature range of 285-310 K, with an

55

Arrhenius expression of k = 3.35 × 10-12 exp[(983 ±130)/T] cm3 molecule-1 s-1.13 That

56

experimental work also identified several products from the OH-initiated oxidation of AcAc,

57

including methylglyoxal (MG), acetic acid (AA), and peroxyacethyl nitrate (PAN).13 On the

58

basis of their measured products, a mechanism for the reaction of AcAc with OH has been

59

postulated, involving the initial OH addition to C2 and C3 positions followed by O2

60

addition.13

61

In contrast to investigation on the industrial applications of AcAc9-12, limited previous

62

work exists on the atmospheric oxidation mechanism of AcAc, hindering accurate assessment

63

of its roles in the formation of ozone and fine particulate matter (PM). For example, MG,

64

organic acids, and PAN have been identified as the critical species leading to SOA

65

formation.14,15,19-24 Specifically, organic acids play important roles in new particle formation

66

and growth25,26 and acid-base reactions27,28, while oligomerization of small α-dicarbonyls

67

represents a major source of SOA on the urban, regional, and global scales.29,30 The

68

atmospheric sources of small α-dicarbonyls, organic acids, and PAN, however, remain

69

poorly quantified.14 The current atmospheric chemical mechanism for the AcAc oxidation has

70

been proposed on the basis of the an environmental chamber experiment,13 which has

4

ACS Paragon Plus Environment

Page 4 of 32

Page 5 of 32

Environmental Science & Technology

71

provided the critical information on the initial kinetic and products of the AcAc oxidation.

72

However, extrapolation of the kinetics and mechanism of the AcAc reactions from the

73

measured product yields is challenging, since a product formation typically involves multiple

74

possible steps and pathways and the product is subject to secondary reactions or photolysis.

75

In addition, there exist additional intricate difficulties using the chamber method. Noticeably,

76

the limitations of the chamber method include a long reaction time, higher reactant

77

concentrations, and wall loss.31,32 In particular, the significance of wall loss for reactive and

78

condensable species using the chamber method has been demonstrated.33-36 Furthermore, no

79

theoretical results are available on the atmospheric chemistry of AcAc.

80

In this work, we have investigated the detailed oxidation mechanism of AcAc with OH

81

employing quantum chemical and kinetic rate calculations within the tropospheric

82

temperature range of 200-310 K. We also present field measurements of AcAc in Nanjing

83

and Beijing, China using proton-transfer reaction mass spectrometry (PTR-MS). The

84

atmospheric fate and oxidation products of AcAc are assessed, and the implications of our

85

results for ozone and SOA formation are discussed.

86

METHODS

87

The electronic structures and energy calculations were carried out with the Gaussian

88

09 program suite.37 Geometrical optimization of all stationary points (SPs), such as the

89

reactants, transition states (TSs), complexes, intermediates, and products, was performed

90

using the M06-2X level with the 6-311G(d,p) basis set denoted as the M06-2X/6-311G(d,p)

5

ACS Paragon Plus Environment

Environmental Science & Technology

91

level. The M06-2X functional is a high-nonlocality functional with double the amount of

92

nonlocal exchange (2X), with reliable performance for the thermochemistry, hydrogen

93

bonding, kinetics, and weak interactions.38 In addition, the MPWB1K/, B3LYP/, MP2/ and

94

QCISD/6-311G(d,p) levels were also employed to optimize the geometry and to validate the

95

convergence of the predicted geometries. The MPWB1K and B3LPY methods and the MP2

96

method represent the classic density functional theory and the classic Ab Initio theory,

97

respectively, while the QCISD method corresponds to a higher electronic correlation method.

98

Frequency calculations were carried out at the M06-2X/6-311G(d,p) level to determine all

99

SPs as a real local minima (without any imaginary frequency) or a TS (with only one

100

imaginary frequency). The evaluation of the vibrational frequencies confirmed that all

101

optimized geometries represented the minima on the potential energy surfaces (Table S1).

102

Intrinsic reaction coordinate (IRC) calculations were performed to confirm the connection

103

between the TSs and their corresponding reactants and products.7,39 The potential energy

104

surface (PES) was further refined by the M06-2X/6-311++G(3df,3pd) level to yield more

105

accurate energetics. Because kinetic calculations of the organic reaction systems were highly

106

sensitive to the predicted energetics, single point energy calculations were performed to

107

refine the PES using the QCISD(T)/6-311+G(2df,p) and CCSD(T)/6-311+G(2df,2p) levels.

108

The dual-level approach was denoted as X//Y, where a single-point energy calculation at

109

level X was carried out for the geometry optimized at a lower level Y. In all cases, the

110

energies were calculated relative to the corresponding reactants including ZPE corrections.

6

ACS Paragon Plus Environment

Page 6 of 32

Page 7 of 32

Environmental Science & Technology

111

∆Ea is defined as the activation energy (i.e., ∆Ea = ETS - Ereactants), while ∆Er is defined as the

112

reaction energy (∆Er = Eproducts - Ereactants). The natural bond orbital (NBO) analysis was

113

carried out to confirm the favorable reaction pathways. All barrierless processes were

114

verified by performing the pointwise potential curve (PPC) scan. Furthermore, the effect of

115

the basis set superposition error on the energies was considered using the counterpoise

116

method described by Boys and Bernardi40 to evaluate the stability of the complexes. All pre-

117

complexes existed, when BSSE correction was included (see Supporting Information). On

118

the basis of the predicted PES, the kinetics calculations, i.e., the rate constants and product

119

distributions, were performed using the Polyrate program41 with the generalized transition-

120

state theory (more details in Supporting In formation).7,31,39 The most stable structure for each

121

pathway was used in the kinetic study.

122

In addition, we measured the concentration of AcAc in Nanjing and Beijing, China on

123

the basis of the proton-transfer reaction,42 i.e., H3O+ + AcAc  H2O + AcAc• H+. The

124

measurements in Nanjing were made using a high-resolution time-of-flight chemical

125

ionization mass spectrometer (HR-ToFCIMS, Aerodyne Res. Inc., USA)43, while the

126

measurements in Beijing were made using a quadruple mass spectrometer.30 Fig. S1 shows

127

the high-resolution fit of the AcAC peak at m/z = 101.06, confirming the detection of AcAc.

128

In this work, the observation sites were located on the campus of Nanjing University of

129

Information Science & Technology (NUIST) in Nanjing and on the campus of Tsinghua

130

University in Beijing.

7

ACS Paragon Plus Environment

Environmental Science & Technology

131

RESULTS AND DISCUSSION

132

Initial reaction of AcAc with OH. There exists an equilibrium between enolic and ketonic

133

isomers of AcAc (Figure S2), with the enolic isomer representing the main form in the gas-

134

phase.44 To systematically assess the OH-initial oxidation of AcAc, we considered both the

135

reactions of enolic and ketonic isomers, denoted by enol-AcAc and keto-AcAc, respectively.

136

The geometries of the two isomers optimized at the M06-2X/6-311G(d,p) level are displayed

137

in Figure 1. For comparison, the geometries and frequencies using other level calculations,

138

including MPWB1K/, B3LYP/, MP2/ and QCISD/6-311G(d,p), are included in Figure S2

139

and Table S1. The structural parameters obtained by the five levels are similar, and the

140

largest discrepancies are within 0.9º in the bond angles and 0.07 Å in the bond lengths. The

141

calculated frequencies at the M06-2X level agree with those obtained by the B3LYP and

142

MP2 levels, with the maximum error of less than 10%. Hence, the M06-2X level theory

143

accurately describes the geometry optimization and vibrational frequency calculations for the

144

AcAc reaction system. Figure 1 shows that enol-AcAc exhibits a conjugated π-electron

145

character that enhances the reactivity of the C atoms. As a result, the C atoms in enol-AcAc

146

are more susceptible to attack by OH than those in keto-AcAc. Figures 1 and S2 depict the

147

optimized geometries of SPs in the OH-AcAc reaction at the M06-2X/6-311G(d,p) level. The

148

absolute energies, ZPEs and Cartesian coordinates are also included in Table S1.

149

The PESs for possible pathways of the OH-initial reaction of enol-AcAc are presented

150

in Figure 2. In addition, the methods of QCISD(T) and CCSD(T) are performed. As

8

ACS Paragon Plus Environment

Page 8 of 32

Page 9 of 32

Environmental Science & Technology

151

discussion in Supporting Information, the M06-2X method is suitable to predict the energies

152

and represents a compromise between the computational accuracy and efficiency.

153

reaction for enol-AcAc with OH occurs via two distinct pathways (Figure 2): H-abstraction

154

from the two methyl groups (e-Rabs1 and e-Rabs2) and OH-addition to either C2 or C3

155

position (e-Radd1 or e-Radd2). For each pathway, a pre-reactive complex is identified prior to

156

the corresponding TS or products. The pre-reactive complexes are consistently more stable

157

than their corresponding reactants (Figure 2). The reaction energies of e-PCadd1 and e-PCadd2

158

are by -5 kcal mol-1 than those of the reactants, verifying that the enol-AcAc oxidation

159

proceeds via the pre-reactive complex and TS prior to the product formation. The structures

160

of the pre-reactive complexes are similar to those of the reactants, except for the forming

161

bond. For instance, the forming C‒O distances are 2.78 and 2.58 Å for the e-PCadd1 and e-

162

PCadd2 (Figure S4a), respectively, while the other bond distances are similar to those of the

163

reactants. The existence of the pre-reactive complexes also impacts the reaction kinetics,45

164

and the OH-AcAc oxidation is expected to exhibit a negative temperature effect.

The

165

Figure 2 shows that the activation energies (∆Ea) of the two H-abstraction pathways (e-

166

Rabs1 and e-Rabs2) are 3.04 and 0.89 kcal mol-1, respectively, which are at least 4 kcal mol-1

167

higher than the two OH-addition pathways (e-Radd1 and e-Radd2). The calculated reaction

168

energy (∆Er) of -28.96 kcal mol-1 for e-Radd1 is by about 11 kcal mol-1 lower than that of e-

169

Radd2. The lower exothermicity for e-Radd1 is explained by its structural characteristics

170

according to the Hammond postulate,46 since this pathway proceeds with an earlier TS due to

9

ACS Paragon Plus Environment

Environmental Science & Technology

171

an elongated C−O distance (Figure 1). However, the ∆Ea value of e-Radd2 is 1.05 kcal mol-1

172

lower than that of e-Radd1 (Figure 2), indicating that e-Radd1 is thermodynamically favored

173

but e-Radd2 is kinetically favored. The natural bond orbital (NBO) charges are 0.496 and

174

0.573 e for the C2 and C3 positions, respectively. Since the more positive potential bond is

175

easily attacked by the nucleophiles (OH), the OH addition to the C3 position (e-Radd2) is

176

more favorable than that to the C2 position (e-Radd1).

177

The reaction of keto-AcAc with OH also occurs via two pathways, i.e., H-abstraction

178

from the methyl and methylene groups (k-Rabs1 and k-Rabs2) and OH-addition to C4 position

179

(k-Radd1). OH addition to keto-AcAc proceeds via the pre-reactive complex prior to the TS,

180

with a ∆Ea value of 6.11 kcal mol-1. The OH-addition to keto-AcAc possesses a higher barrier

181

(by about 10 kcal mol-1) than those of the H-abstraction pathways; the distinct behaviors

182

between OH additions to enol- and keto-AcAc are explained because of the conjugative

183

effect.

184

The occurrence of the H-abstraction TSs is further evaluated using the L parameter

185

(L= δ(H-O)) according to our previous study.47 This parameter not only quantifies whether the

186

TS structure exhibits a product-like (L > 1) or reactant-like (L < 1) character but also reflects

187

whether the pathway is exothermic or endothermic. The L value of 0.34 for k-TSabs1 is two

188

times higher than that of k-TSabs2 (Figure 1). Hence, the k-Rabs2 pathway with an earlier TS

189

corresponds to a larger ∆Er. As shown in Figure 2, the ∆Er of k-Rabs2 is -26.57 kcal mol-1,

190

which is by about 4.41 kcal mol-1 more negative than that of k-Rabs1. On the other hand, the

δ(C-H)

10

ACS Paragon Plus Environment

Page 10 of 32

Page 11 of 32

Environmental Science & Technology

191

k-Rabs1 pathway corresponds to a ∆Ea value of 0.56 kcal mol-1, and there exists a hydrogen

192

bond with the distance of 2.02 Å in the TS (Figure 1). In order to further assess the

193

occurrence for H-abstraction, the dissociation energies (D0298(C‒H)) of methylene (-CH2-)

194

and methyl (-CH3) groups are calculated at the M06-2X//M06-2X level. The (D0298 (C‒H))

195

values are 89.77 kcal mol-1 for the -CH2- group and 94.14 kcal mol-1 for the -CH3 group,

196

respectively, indicating that the H atom in the -CH2- group (k-Rabs2) are more reactive than

197

that in the-CH3 group (k-Rabs1). Considering the ∆Ea and ∆Er values, the reactivity of the H

198

atom in the two groups is dominantly affected by the steric effect rather than its stability,

199

attributable to the presence of the -CH2- group in the middle of AcAc.

200

The rate constants of the OH-initial reaction pathways of keto- and enol-AcAc are

201

calculated and summarized in Table S2, and the temperature dependences of the branching

202

ratios (Γ) are shown in Figure S5. For the reaction of OH with enol-AcAc, the rate constants

203

of e-Radd1 and e-Radd2 pathways are 3.78 × 10-11 and 7.46 × 10-11 cm3 molecule-1 s-1 at 298 K,

204

respectively, which are by 2‒3 orders of magnitude higher than those of the two H-

205

abstraction pathways (e-Rabs1 and e-Rabs2) (Table S2). The contribution of the combined H-

206

abstraction pathways to the total rate constant is less than 1%, suggesting that the H-

207

abstraction pathway is of minor importance. As shown in Figure S5a, the Γ of e-Radd2 is

208

greater than 66% in the temperature range of 237‒298 K. Hence, OH addition at C3 position

209

(e-Add-2) is more favorable, while the Γ value for OH addition at C2 position (e-Add-1) is

210

33% at 298 K.

11

ACS Paragon Plus Environment

Environmental Science & Technology

211

The rate constants of the OH-addition pathway for keto-AcAc contributes negligibly to

212

the total rate constants (Γ=0), and the sum Γ of the H-abstraction pathway equals to 1, in

213

contrast to the case of enol-AcAc (Figure S4b). For example, the rate constants of k-Rabs1 and

214

k-Rabs2 are 4.31 × 10-13 and 5.86 × 10-14 cm3 molecule-1 s-1 at 298 K, respectively,

215

contributing 88% and 12% to the rate constant for keto-AcAc. Our results of dominant H-

216

abstraction for k-Rabs1 are consistent with those obtained by Holloway et al16 but are in

217

contrast with those by Zhou et al.13 The reactivity of the -CH2- group is likely overestimated

218

using the structure-activity relationship, because the impact of the steric effect is not

219

considered.13

220

The calculated total rate constants (i.e., ktotal, the sum of calculated rate constants for all

221

pathways) for the OH-AcAc reaction are presented in Figure 3, along with comparison with

222

the available experimental data. Our derived Arrhenius expression is ktotal = 1.69 × 10-13

223

exp(1935/T) cm3 molecule-1 s-1 over the temperature range of 200-310 K. The rate constant

224

for the keto-pathway (kketo) is much smaller than that of the enol-pathway (kenol).

225

Furthermore, the previous studies have reported AcAc in gas phase exists predominantly in

226

the enol-form at room temperature.13,52 Hence, OH addition to enol-AcAc is favorable than

227

addition to keto-AcAc. As is evident from Figure 3, our results at the M06-2X//M06-2X level

228

also compare favorably with the available experimental data, considering the respective

229

uncertainties. For example, the total rate constant at 298 K is calculated to be 11.1 × 10-11

230

cm3 molecule-1 s-1, consistent with the rate constant of (9.05 ± 1.81) × 10-11 cm3 molecule-1 s-1

12

ACS Paragon Plus Environment

Page 12 of 32

Page 13 of 32

Environmental Science & Technology

231

measured by Zhou et al.13 but somewhat higher than those reported by Bell et al.17 and

232

Holloway et al.16 Our calculations show a strong negative temperature-dependence for the

233

rate constants in the temperature range of 200-310 K, which is attributable to the presence of

234

the pre-reactive complex. Also for comparison, Bell et al.17 reported report an activation

235

parameter (Ea/R) of 1260 K, comparable to our theoretical value. Clearly, the M06-2X

236

method provides a reliable description for the initial kinetics of the atmospheric oxidation of

237

AcAc.

238

Subsequent oxidation of the OH-AcAc adduct. The attack of e-Add-1 and e-Add-2 by O2

239

yields two peroxy radicals (RO2-1 and RO2-2), which further undergo NO-addition to form

240

peroxy nitrites (RO2NO-1 and RO2NO-2) followed by NO2-elimination to form alkoxy

241

radicals (RO1 and RO2). Such multistep processes are described as the following (Figure 4), R11 or R21: + O2

R12 or R22: +NO

242

e-Add-1 or e-Add-2  RO2-1 or RO2-2  RO2NO-1 or

243

RO2NO-2  RO-1 or RO-2

R13 or R23: -NO2

244

The vibrational frequencies, absolute energies, ZPEs and Cartesian coordinates of the

245

relevant species are included in Table S3. As shown in Figure 4, the ∆Ea of pathway R11 is

246

21.33 kcal mol-1, but the TS of the association of e-Add-2 with O2 (R21) is not identified.

247

The PPC was performed to confirm a barrierless process for R21 (Figure S5). The ∆Er values

248

of R11 and R21 are -26.24 and -36.98 kcal mol-1, respectively. The large ∆Ea and instability

249

of RO2-1 indicate that its formation is thermodynamically and kinetically less favorable. To

250

assess the competition between decomposition and O2 addition for e-Add-1, we calculated

251

the rate constants for both pathways, with the values of 4.80 × 10-7 s-1 and 7.75 × 10-26 cm3

252

molecule-1 s-1, respectively. Hence, decomposition of e-Add-1 to enol-AcAc and OH is more 13

ACS Paragon Plus Environment

Environmental Science & Technology

253

favorable than the combination with O2 (with an equivalent first-order rate constant of 3.81 x

254

10-7 s-1). Considering the branching ratios between e-Add-1 and e-Add-2, it is estimated that

255

about 15% of e-Add-1 reacts with O2 to further propagate the oxidation (Figure 5).

256

The association reactions of RO2-1 and RO2-2 with NO (R12 and R22) are barrierless

257

and exothermic, with the ∆Er values of -22.42 and -21.12 kcal mol-1, respectively. The

258

produced peroxy nitrites (RO2NO-1 and RO2NO-2) further undergo NO-elimination via a

259

barrierless process to form the alkoxy radicals (RO-1 and RO-2), with the ∆Er values of

260

13.88 and 9.44 kcal mol-1, respectively. According to previous studies,48,49 there are three

261

primary reactions for the alkoxy radicals, i.e., dissociation, isomerization, and reaction with

262

O2. The reaction of with O2 is competitive only if isomerization is impossible and

263

dissociation forms primary alkyl radicals. The isomerization occurs only if there exists a H

264

atom located at four carbons away from the radical center (which is absent in our reaction

265

system. Hence, we focus on the subsequent dissociation of RO-1 and RO-2 involves the

266

following stepwise processes,  



267

RO-1  TS12 or TS13  ER and MG or Acetyl Radical and PD

268

RO-2  TS22 or TS23  MR and AA or butane-2,4-dione-3,4-diol

269

and CH3

   



270

Decomposition of RO-1 via TS12 or TS13 yields (R14-1) MG and ethanediol radical (ER)

271

or (R14-2) acetyl radical (AR) and propanal-2,2-diol (PD), with the ∆Ea values of 0.40 and

272

3.73 kcal mol-1, respectively. The small ∆Ea difference between R14-1 and R14-2 indicates

273

that the products of the two pathways are both accessible. The ∆Ea values of the two

274

decomposition pathways of RO-2 are 1.48 and 8.64 kcal mol-1 to form (R24-1) AA and

275

methylglyoxal radical (MR) and (R24-2) butane-2,4-dione-3,4-diol and methyl radical (CH3), 14

ACS Paragon Plus Environment

Page 14 of 32

Page 15 of 32

Environmental Science & Technology

276

respectively. The large ∆Ea and small ∆Er for R24-2 indicate a minor importance to form

277

butane-2,4-dione-3,4-diol and CH3. Although the transition states of R14 and R24-2

278

pathways have lower energies than those of the corresponding products (Figure 4), there exist

279

the product complexes (COM12 and COM13) at the exit channels. Hence, AA and MG are the

280

major products from both R14-1 and R24-1.

281

Subsequently, ER, PD, and MR further react with O2 via the following stepwise processes, R15: + O2

R17

-HO2

R16: + O2

R18: + NO2

282

ER  ER-RO2  TS15  AA

283

AR  AR-RO2  PAN

284

MR  TS23  MR-RO2  TS24  MG

285

The ∆Er values of R15 and R16 are -41.94 and -35.53 kcal mol-1, respectively, via the

286

barrierles processes to form two peroxy radicals (ER-RO2 and AR-RO2). The subsequent

287

reaction of AR-RO2 with NO2 is also barrierless, with a ∆Er value of 24.76 kcal mol-1 to yield

288

PAN. The ∆Ea value of ER-RO2 decomposition to form AA is 6.59 kcal mol-1. The pathway

289

from MR to MG corresponds to successively endothermic processes (R25 and R26) but

290

occurs promptly, considering the large exothermicity for the formation of e-Add-2 and

291

RO2NO-2.

R25: + O2

R26

-HO2

292

It is plausible that there are additional pathways to those investigated in our present

293

work, which require further experimental and theoretical studies. Our results indicate that

294

MG and AA represent the most favorable products from the oxidation of AcAc initiated by

295

OH, both arising from the pathways via e-Add-1 and e-Add-2. For comparison, the previous

15

ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 32

296

experimental measurements by Zhou et al. obtained the yields of (20.8 ± 4.5)% and (16.9 ±

297

3.4)% for MG and AA, respectively. In addition, we predict a minor pathway via e-Add-1 to

298

form PAN, consistent with the work by Zhao et al. for a small yield for the formation of PAN

299

(about 2%).

300

Atmospheric lifetime of AcAc. The lifetime of VOCs is a key parameter to assess their roles

301

in the formation of ozone and SOA. The OH initiated oxidation represents the dominant

302

daytime mechanism in regulating the lifetimes of VOCs.1,50 We evaluate the atmospheric

303

lifetime (τ) for AcAc at the different altitude and [OH] according to = 

304

[OH] and ktotal are the OH concentration and total rate constant, respectively. Using a lapse

305

rate of 6.5 K km-1 for the typical tropospheric condition,47 we determine the lifetime of AcAc,

306

and results are presented in Table S4. At the ground level with [OH] = 1 × 106 molecules cm-

307

3

308

rapidly with increasing height and [OH]. For example, the τ value decreases to 0.16 hr as

309

[OH] increases to 1.5 × 107 molecules cm-3.

310

ATMOSPHERIC IMPLICATIONS

  []

, where

(12 hr daytime average), the lifetime of AcAc is 1.54 hr. The AcAc lifetime decreases

311

Ketones are generally less reactive because the less reactive keto-forms are generally

312

preferred than the more reactive enol-forms,1 while for AcAc the enol-form is dominant due

313

to intramolecular hydrogen bonding.44 We have assessed the mechanism, kinetics, and

314

atmospheric fate of the OH-initiated oxidation for the enolic and ketonic isomers of AcAc

315

using quantum chemical and kinetic rate calculations (Figure 5). OH addition to enol-AcAc is

16

ACS Paragon Plus Environment

Page 17 of 32

Environmental Science & Technology

316

more favorable than addition to keto-AcAc, because of the conjugated π-electron effect.

317

Consequently, the rate constant for the keto-pathway is much smaller than that of the enol-

318

pathway, and the reaction of the enol-AcAc with OH represents the dominant pathway for

319

AcAc oxidation in the troposphere, largely determining the fate and impacts of its products.

320

For the reaction of the enol-AcAc with OH, the rate constants for OH addition are by 2‒3

321

orders of magnitude higher than those for H-abstraction, and OH addition to enol-AcAc

322

occurs via a prereactive complex. Our derived total rate constant for the reaction of AcAc

323

with OH is ktotal = 1.69 × 10-13 exp(1935/T) cm3 molecule-1 s-1 over the temperature range of

324

200-310 K and has a value of 11.1 × 10-11 cm3 molecule-1 s-1 at 298 K, consistent with the

325

previous experimental measurements.13,16,17

326

For the subsequent reactions of the OH-enol-AcAc adduct, O2 addition to the C3-adduct

327

proceeds barrierlessly with a large negative reaction energy (e-add-2), while O2 addition to

328

the C2-adduct occurs with a high activation barrier (about 21 kcal mol-1) and a smaller

329

exothermicity (e-add-1). AA and MG are the most favorable products from the two peroxy

330

radical pathways (RO2-1 and RO2-2), with a comparable yield. The branching ratio of about

331

81% leading to the formation of the two peroxy radicals likely corresponds to the upper limit

332

for the AA and MG yields, indicating the dominant production of the two species from the

333

OH-initiated oxidation of AcAc. The previous experimental study for the OH-initiated

334

oxidation of AcAc by Zhou et al. obtained the yields for AA (about 17%) and MG (about

335

21%).13 Considering possible wall loss and secondary reactions (including photolysis for

17

ACS Paragon Plus Environment

Environmental Science & Technology

336

MG) of these species in the chamber work,33-36 those measured yields likely correspond to

337

the lower experimental limits. On the other hand, our results indicate a minor pathway

338

leading to PAN, consistent with the small experimental yield of PAN (2%).13 Other plausible

339

products from the OH-initiated oxidation of AcAc include organic nitrates (RONO2) directly

340

arising from the peroxy nitrites (RO2NO-1 and RO2NO-2) and minor butane-2,4-dione-3,4-

341

diol. However, direct formation of organic nitrates from peroxy nitrites typically constitutes a

342

minor pathway, in contrast to the dominant formation for alkoxy radicals.51 Using the

343

predicted temperature-dependence kinetic data, we estimate a lifetime of less than a few

344

hours due the OH-initiated oxidation of AcAc throughout the entire tropospheric conditions.

345

Figure 6 shows a time series of AcAc measurements in Nanjing and Beijing, China. The

346

measured AcAc concentration varies from ppt to ppb levels and exhibits a diurnal variation in

347

both locations, which is likely regulated by the emission, planetary boundary layer height,

348

and photochemical activity. Such atmospheric concentrations of AcAc are clearly significant,

349

considering its high reactivity with OH and short atmospheric lifetimes under tropospheric

350

conditions. Since AA and MG are important SOA precursors,14,15,19-24,52 accurate

351

determination of their yields from AcAc oxidation by OH is critical to assess SOA formation

352

under polluted environments. Furthermore, organic peroxy radical (RO2) and hydroperoxy

353

radical (HO2) are produced from the OH-initiated oxidation of enol-AcAc (Figure 4),

354

facilitating the cycling between NO and NO2 and producing tropospheric ozone.51 Hence,

355

because its high reactivity and dominant yields for MG and AA, the photochemical oxidation

18

ACS Paragon Plus Environment

Page 18 of 32

Page 19 of 32

Environmental Science & Technology

356

of AcAc may contribute importantly to ozone and SOA formation under polluted

357

environments, with implications for air quality, human health, and climate. Our results

358

provide the kinetic and mechanistic data for inclusion of the oxidation of AcAc in

359

atmospheric models. Future studies are necessary to assess the impacts of AcAc on ozone

360

and SOA formation using chemical transport models, with the consideration of its emission

361

inventory, chemistry, and transport.

362

ASSOCIATED CONTENT

363

Supporting Information

364

The structures, Cartesian coordinates, frequencies, zero-point energies, and absolute energies

365

of all relevant species involved in the title reaction, along with the branching ratios, the rate

366

constants, and the lifetimes are listed in the Supporting Information. This material is

367

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

368

AUTHOR INFORMATION

369

Corresponding Authors

370

* Phone: 86-20-23883536; Fax: 86-20-23883536; E-mail address: [email protected]

371

* Phone: 979-845-7656; Fax: 979-862-4466; E-mail address: [email protected]

372

Notes

373

The authors declare no competing financial interest.

374

ACKNOWLEDGMENTS

375

This work was financially supported by National Natural Science Foundation of China

376

(41675122, 41425015, U1401245 and 41373102), Science and Technology Program of

19

ACS Paragon Plus Environment

Environmental Science & Technology

377

Guangzhou City (201707010188), Team Project from the Natural Science Foundation of

378

Guangdong Province, China (S2012030006604), and the Special Program for Applied

379

Research on Super Computation of the NSFC-Guangdong Joint Fund (the second phase), and

380

the National Supercomputing Centre in Guangzhou (NSCC-GZ). R.Z. acknowledged the

381

support from the Robert A. Welch foundation (A-1417).

382

REFERENCES

383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411

(1) Atkinson, R.; Arey, J. Atmospheric degradation of volatile organic compounds. Chem. Rev. 2003, 103 (12) 4605- 4638. (2) Dalmasso, P. R.; Taccone, R. A.; Nieto, J. D.; Cometto, P. M.; Cobos, C. J.; Lane, S. I. Reactivity of hydrohaloethers with OH radicals and chlorine atoms: Correlation with molecular properties. Atmos. Environ. 2014, 91, 104- 109. (3) Lewis, A. C.; Carslaw, N.; Marriott, P. J.; Kinghorn, R. M.; Morrison, P.; Lee, A. L.; Bartle, K. D.; Pilling, M. J. A larger pool of ozone-forming carbon compounds in urban atmospheres. Nature 2000, 405 (6788) 778-781. (4) Fares, S.; Paoletti, E.; Loreto, F.; Brilli, F. Bidirectional flux of methyl vinyl ketone and methacrolein in trees with different isoprenoid emission under realistic ambient concentrations. Environ. Sci. Technol. 2015, 49 (13) 7735- 7742. (5) Liu, Y.; Shao, M.; Kuster, W. C.; Goldan, P. D.; Li, X.; Lu, S.; Gouw, J. A. D. Source identification of reactive hydrocarbons and oxygenated VOCs in the summertime in Beijing. Environ. Sci. Technol. 2008, 43 (1) 75- 81. (6) Arnold, F.; Bürger, V.; Droste‐Fanke, B.; Grimm, F.; Krieger, A.; Schneider, J.; Stilp, T. Acetone in the upper troposphere and lower stratosphere: Impact on trace gases and aerosols. Geophys. Res. Lett. 1997, 24 (23) 3017- 3020. (7) Ji, Y.; Wang, L.; Li, Z.; Liu, J.; Sun, C. Theoretical dynamic studies on the reaction of CH3C (O) CH3− nFn with the hydroxyl radical and the chlorine atom. ChemPhysChem 2006, 7 (8) 1741- 1749. (8) Atkinson, R.; Tuazon, E. C.; Aschmann, S. M. Atmospheric chemistry of 2-pentanone and 2-heptanone. Environ. Sci. Technol. 2000, 34 (4) 623- 631. (9) Warneke, J.; Van Dorp, W. F.; Rudolf, P.; Stano, M.; Papp, P.; Matejčík, Š.; Borrmann, T.; Swiderek, P. Acetone and the precursor ligand acetylacetone: Distinctly different electron beam induced decomposition? Phys. Chem. Chem. Phys. 2014, 17 (2) 12041216. (10) Ho, Y. S.; Wu, C. H.; Chen, Y. S.; Liu, C. K. Effects of molar ratio of acetylacetone to aluminum precursor on the yttrium aluminum garnet (Y3Al5O12, YAG) formation by a sol–gel process. Int. J. Appl. Ceram. Technol. 2015, 12 (S2), E53–E58. 20

ACS Paragon Plus Environment

Page 20 of 32

Page 21 of 32

412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452

Environmental Science & Technology

(11) Yang, H.; Sun, H.; Zhang, S.; Wu, B.; Pan, B. Potential of acetylacetone as a mediator for Trametes versicolor laccase in enzymatic transformation of organic pollutants. Environ. Sci. Pollut. Res. 2015, 22 (14) 10882-10889. (12) El-Sayed, H. A.; Said, S. A.; Amr, A. E.-G. E. Synthesis of some fused heterocyclic systems and their nucleoside candidates. Res. Chem. Intermed. 2014, 40 (2) 833- 845. (13) Zhou, S.; Barnes, I.; Zhu, T.; Bejan, I.; Albu, M.; Benter, T. Atmospheric chemistry of acetylacetone. Environ. Sci. Technol. 2008, 42 (21) 7905- 7910. (14) Zhang, R.; Wang, G.; Guo, S.; Zamora, M. L.; Ying, Q.; Lin, Y.; Wang, W.; Hu, M.; Wang, Y. Formation of urban fine particulate matter. Chem. Rev. 2015, 115 (10) 38033855. (15) Guo, S.; Hu, M.; Zamora, M. L.; Peng, J.; Shang, D.; Zheng, J.; Du, Z.; Wu, Z.; Shao, M.; Zeng, L. Elucidating severe urban haze formation in China. Proc. Natl. Acad. Sci. U. S. A. 2014, 111 (49) 17373- 17378. (16) Holloway, A.-L.; Treacy, J.; Sidebottom, H.; Mellouki, A.; Daële, V.; Le Bras, G.; Barnes, I. Rate coefficients for the reactions of OH radicals with the keto/enol tautomers of 2, 4-pentanedione and 3-methyl-2, 4-pentanedione, allyl alcohol and methyl vinyl ketone using the enols and methyl nitrite as photolytic sources of OH. J. Photochem. Photobiol., A 2005, 176 (1) 183- 190. (17) Bell, P.; Nicovich, J. M.; Wine, P. H. Acetylacetone photolysis at 248 nm: hydroxyl radical yield and temperature-dependent rate coefficients for the OH + acetylacetone reaction. Proceedings of 19th International Symposium on Gas Kinetics, Orleans, France, July 22-27, 2006; pp 79-80. (18) Dagaut, P.; Wallington, T. J.; Liu, R.; Kurylo, M. J. A kinetic investigation of the gasphase reactions of hydroxyl radicals with cyclic ketones and diones: mechanistic insights. J. Phys. Chem. A. 1988, 92 (15) 4375- 4377. (19) Tan, Y.; Carlton, A. G.; Seitzinger, S. P.; Turpin, B. J., SOA from methylglyoxal in clouds and wet aerosols: Measurement and prediction of key products. Atmos. Environ. 2010, 44 (39), 5218-5226. (20) Tan, Y.; Lim, Y. B.; Altieri, K. E.; Seitzinger, S. P.; Turpin, B. J., Mechanisms leading to oligomers and SOA through aqueous photooxidation: insights from OH radical oxidation of acetic acid and methylglyoxal. Atmos. Chem. Phys. 2012, 11 (11) 1831918347. (21) Lim, Y. B.; Tan, Y.; Turpin, B. J., Chemical insights, explicit chemistry and yields of secondary organic aerosol from methylglyoxal and glyoxal. Atmos. Chem. Phys. 2013, 13 (17) 4687-4725. (22) Lim, Y.B., Tan, Y., Perri, M.J., Seitzinger, S.P. & Turpin, B.J. Aqueous chemistry and its role in secondary organic aerosol (SOA) formation. Atmos. Chem. Phys. 2010, 10, 10521–10539. (23) De Haan, D.O. et al. Brown Carbon Production in Ammonium- or Amine-Containing Aerosol Particles by Reactive Uptake of Methylglyoxal and Photolytic Cloud Cycling. Environ. Sci. Technol. 2017, 51, 7458–7466. 21

ACS Paragon Plus Environment

Environmental Science & Technology

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

(24) De Haan, D. O. et al. Nitrogen-containing, light-absorbing oligomers produced in aerosol particles exposed to methylglyoxal, photolysis, and cloud cycling. Environ Sci Technol, 2018, 52, 4061-4071. (25) Zhang, R.; Suh, I.; Zhao, J.; Zhang, D.; Fortner, E. C.; Tie, X.; Molina, L. T.; Molina, M. J. Atmospheric new particle formation enhanced by organic acids. Science 2004, 304 (5676) 1487- 1490. (26) Altieri, K. E.; Seitzinger, S. P.; Carlton, A. G.; Turpin, B. J.; Klein, G. C.; Marshall, A. G. Oligomers formed through in-cloud methylglyoxal reactions: Chemical composition, properties, and mechanisms investigated by ultra-high resolution FT-ICR mass spectrometry. Atmos. Environ., 2008, 42(7) 1476-1490. (27) Qiu, C.; Zhang, R. Multiphase chemistry of atmospheric amines. Phys. Chem. Chem. Phys. 2013, 15 (16) 5738- 5752. (28) Qiu, C.; Wang, L.; Lal, V.; Khalizov, A. F.; Zhang, R. Heterogeneous reactions of alkylamines with ammonium sulfate and ammonium bisulfate. Environ. Sci. Technol. 2011, 45 (11) 4748- 4755. (29) Gomez, M. E.; Lin, Y.; Guo, S.; Zhang, R. Heterogeneous chemistry of glyoxal on acidic solutions an oligomerization pathway for secondary organic aerosol formation. J. Phys. Chem. A. 2014, 119 (19) 4457- 4463. (30) Zhao, J.; Levitt, N. P.; Zhang, R.; Chen, J. Heterogeneous reactions of methylglyoxal in acidic media: Implications for secondary organic aerosol formation. Environ. Sci. Technol. 2006, 40 (24) 7682- 7687. (31) Ji, Y.; Li, Y.; An, T.; Zhang, R. Reply to Newland et al.: The dominant phenolic pathway for atmospheric toluene oxidation. Proc. Natl. Acad. Sci. U. S. A. 2017, 114 (38) e7858e7859. (32) Ji, Y. et al. Reassessing the atmospheric oxidation mechanism of toluene. Proc. Natl. Acad. Sci. U. S. A. 2017, 8169- 8174. (33) Zhang, X.; Cappa, C. D.; Jathar, S. H.; McVay, R. C.; Ensberg, J. J.; Kleeman, M. J.; Seinfeld, J. H. Influence of vapor wall loss in laboratory chambers on yields of secondary organic aerosol. Proc. Natl. Acad. Sci. U. S. A. 2014, 111 (16) 5802- 5807. (34) Matsunaga, A.; Ziemann, P. J., Gas-Wall Partitioning of Organic Compounds in a Teflon Film Chamber and Potential Effects on Reaction Product and Aerosol Yield Measurements. Aerosol Sci. Technol. 2010, 44 (10) 881-892. (35) La, Y. S.; Camredon, M.; Ziemann, P. J.; Valorso, R.; Matsunaga, A.; Lannuque, V.; Lee-Taylor, J.; Hodzic, A.; Madronich, S.; Aumont, B., Impact of chamber wall loss of gaseous organic compounds on secondary organic aerosol formation: explicit modeling of SOA formation from alkane and alkene oxidation. Atmos. Chem. Phys. 2016, 16 (3) 1417-1431. (36) Krechmer, J. E.; Pagonis, D.; Ziemann, P. J.; Jimenez, J. L., Quantification of Gas-Wall Partitioning in Teflon Environmental Chambers Using Rapid Bursts of Low-Volatility Oxidized Species Generated in Situ. Environ. Sci Technol. 2016, 50 (11) 5757-5765.

22

ACS Paragon Plus Environment

Page 22 of 32

Page 23 of 32

493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534

Environmental Science & Technology

(37) Frisch, M.; Trucks, G.; Schlegel, H. B.; Scuseria, G.; Robb, M.; Cheeseman, J.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G., Gaussian 09, revision D. 01; Gaussian Inc.: Wallingford CT, 2009. (38) Zhao, Y.; Truhlar, D. G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120 (1-3) 215- 241. (39) Zhang, D.; Zhang, R. Mechanism of OH formation from ozonolysis of isoprene: A quantum-chemical study. J. Am. Chem. Soc. 2002, 124, 2692-2703. (40) Boys, S. F.; Bernardi, F. The calculation of small molecular interactions by the differences of separate total energies. Some procedures with reduced errors. Mol. Phy. 1970, 19, 553-566. (41) Zheng, J., Zhang, S., Lynch, B., Corchado, J., Chuang, Y., Fast, P., Hu, W., Liu, Y.-P., Lynch, G., Nguyen, K. POLYRATE, version 2010-A; Software Manager; University of Minnesota: Minneapolis, MN, 2010. (42) Zhao, J.; Zhang, R. Y., Proton transfer reaction rate constants between hydronium ion (H3O+) and volatile organic compounds. Atmos. Environ. 2004, 38, (14), 2177-2185. (43) Zheng, J.; Ma, Y.; Chen, M.; Zhang, Q.; Wang, L.; Khalizov, A. F.; Yao, L.; Wang, Z.; Wang, X.; Chen, L., Measurement of atmospheric amines and ammonia using the high resolution time-of-flight chemical ionization mass spectrometry. Atmos. Environ. 2015, 102, (0), 249-259. (44) Nakanishi, H.; Morita, H.; Nagakura, S. Electronic structures and spectra of the keto and enol forms of acetylacetone. Bull. Chem. Soc. Jpn. 1977, 50 (9) 2255-2261. (45) Alvarez-Idaboy, J. R.; Mora-Diez, N.; Boyd, R. J.; Vivier-Bunge, A. On the importance of prereactive complexes in molecule-radical reactions: Hydrogen abstraction from aldehydes by OH. J. Am. Chem. Soc. 2001, 123 (9) 2018- 2024. (46) Hammond, G. S. A correlation of reaction rates J. Am. Chem. Soc. 1955, 77 (2) 334-338. (47) Ji, Y.; Wang, H.; Gao, Y.; Li, G.; An, T. A theoretical model on the formation mechanism and kinetics of highly toxic air pollutants from halogenated formaldehydes reacted with halogen atoms. Atmos. Chem. Phys. 2013, 13 (22) 11277-11286. (48) Atkinson, R. Rate constants for the atmospheric reactions of alkoxy radicals: an updated estimation method. Atmos. Environ. 2007, 41, 8468-8485. (49) Kroll, J. H.; Seinfeld, J. H. Chemistry of secondary organic aerosol: Formation and evolution of low-volatility organics in the atmosphere. Atmos. Environ. 2008, 42, 35933624. (50) Wang, H.; Ji, Y.; Gao, Y.; Li, G.; An, T. Theoretical model on the formation possibility of secondary organic aerosol from OH initialed oxidation reaction of styrene in the presence of O2/NO. Atmos. Environ. 2015, 101, 1-9. (51) Zhang, D.; Zhang, R.; Park, J.; North, S. W. Hydroxy peroxy nitrites and nitrates from OH initiated reactions of isoprene J. Am. Chem. Soc. 2002, 124 (32) 9600- 9605. (52) Zhang, R.; Khalizov, A.F.; Wang, L.; Hu, M.; Xu, W. Nucleation and growth of nanoparticles in the atmosphere. Chem. Rev. 2012, 112, 1957-2011. 23

ACS Paragon Plus Environment

Environmental Science & Technology

535 536 537 538 539 540

(53) Zhang, R.; Lei, W.; Tie, X.; Hess, P. Industrial emissions cause extreme urban ozone diurnal variability Proc. Natl. Acad. Sci. U. S. A. 2004, 101 (17) 6346- 6350. (54) Iglesias, E. Determination of the keto-enol equilibrium constants and the kinetic study of the nitrosation reaction of β-dicarbonyl compounds. J. Chem. Soc., Perkin Trans. 2 1997, 431–439.

541 542

24

ACS Paragon Plus Environment

Page 24 of 32

Page 25 of 32

Environmental Science & Technology

543

Figure captions

544

Figure 1. The optimized geometries of the key stationary points of OH-AcAc reaction at the

545 546 547

M06-2X/6-311G(d,p). The bond length is in Å. Figure 2. PES for the OH-initiated reactions of keto- and enol-AcAc (in the unit of kcal mol1

).

548

Figure 3. The rate constants (cm3 molecule-1 s-1) of keto- and enol-AcAc with OH radical

549

against the temperature. The experimental results (i.e., Expt. 1, 2 and 3) are from Zhou

550

et al., Bell et al., and Holloway et al., respectively.

551

Figure 4. PES for (a) the subsequent pathways of e-Add-2 and (b) the competing

552

decomposition and combination with O2 for e-Add-1. The number denotes the value of

553

∆Ea and ∆Er for each reaction step. (unit: kcal mol-1 for energies and cm3 molecule-1 s-1

554

for the rate constants).

555 556 557 558

Figure 5. Schematic representation of the preferred pathways of the OH-AcAc reactions leading to formation of methylglyoxal and acetic acid. Figure 6. Time series of AcAc measured using the PRT- MS method in Nanjing (a) and Beijing (b), China. The major tick mark on the x-axis labels the local time at midnight.

559

25

ACS Paragon Plus Environment

Environmental Science & Technology

560

Figure 1

561

26

ACS Paragon Plus Environment

Page 26 of 32

Page 27 of 32

Environmental Science & Technology

Figure 2

27

ACS Paragon Plus Environment

Environmental Science & Technology

Figure 3

28

ACS Paragon Plus Environment

Page 28 of 32

Page 29 of 32

Environmental Science & Technology

Figure 4

29

ACS Paragon Plus Environment

Environmental Science & Technology

Figure 5

ACS Paragon Plus Environment

Page 30 of 32

Page 31 of 32

Environmental Science & Technology

Figure 6

ACS Paragon Plus Environment

Environmental Science & Technology

TOC

ACS Paragon Plus Environment

Page 32 of 32