Isomerization of Second-Generation Isoprene Peroxy Radicals

Apr 7, 2017 - This finding is consistent with quantum chemical calculations that suggest a net forward rate constant of 0.3–0.9 s–1. Furthermore, ...
0 downloads 0 Views 1MB Size
Subscriber access provided by UB + Fachbibliothek Chemie | (FU-Bibliothekssystem)

Article

Isomerization of second generation isoprene peroxy radicals: epoxide formation and implications for secondary organic aerosol yields Emma L. D'Ambro, Kristian H. Møller, Felipe D. Lopez-Hilfiker, Siegfried Schobesberger, Jiumeng Liu, John E. Shilling, Ben Hwan Lee, Henrik G. Kjaergaard, and Joel A. Thornton Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b00460 • Publication Date (Web): 07 Apr 2017 Downloaded from http://pubs.acs.org on April 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.

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

Page 1 of 30

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

Environmental Science & Technology

Isomerization of second generation isoprene peroxy radicals: epoxide formation and implications for secondary organic aerosol yields Emma L. D’Ambro1,2, Kristian H. Møller3, Felipe D. Lopez-Hilfiker1,a , Siegfried Schobesberger1,b , Jiumeng Liu4, John E. Shilling4,5, Ben Hwan Lee1, Henrik G. Kjaergaard3, Joel A. Thornton*,1,2 1

Department of Atmospheric Sciences, University of Washington, Seattle, Washington 98195, United States 2 Department of Chemistry, University of Washington, Seattle, Washington 98195, United States 3 Department of Chemistry, University of Copenhagen, Universitetsparken 5, DK-2100 Copenhagen Ø, Denmark 4 Atmospheric Sciences and Global Change Division, Pacific Northwest National Laboratory, Richland, WA, 99352, USA 5 Environmental Molecular Sciences Laboratory, Pacific Northwest National Laboratory, Richland, WA, 99352, USA a Now at: Laboratory of Atmospheric Chemistry, Paul Scherrer Institute, Zurich, Switzerland b Now at: Department of Applied Physics, University of Eastern Finland, Kuopio, Finland Author Information Corresponding Author * Phone: 206-543-4010; Email: [email protected]; Mail: 408 ATG Building, Box 351640, University of Washington, Seattle, WA 98195

1 ACS Paragon Plus Environment

Environmental Science & Technology

47

Page 2 of 30

Abstract

48

We report chamber measurements of secondary organic aerosol (SOA) formation from

49

isoprene photochemical oxidation, where radical concentrations were systematically varied and

50

the molecular composition of semi to low volatility gases and SOA were measured online. Using

51

a detailed chemical kinetics box model, we find that to explain the behavior of low volatility

52

products and SOA mass yields relative to input H2O2 concentrations, the second generation

53

dihydroxy hydroperoxy peroxy radical (C5H11O6•) must undergo an intra-molecular H-shift with

54

a net forward rate constant of order 0.1 s-1 or higher. This finding is consistent with quantum

55

chemical calculations which suggest a net forward rate constant of 0.3-0.9 s-1. Furthermore, these

56

calculations suggest the dominant product of this isomerization is a dihydroxy hydroperoxy

57

epoxide (C5H10O5) which is expected to have a saturation vapor pressure ~2 orders of magnitude

58

higher, as determined by group-contribution calculations, than the dihydroxy dihydroperoxide,

59

ISOP(OOH)2 (C5H12O6), a major product of the peroxy radical reacting with HO2. These results

60

provide strong constraints on the likely volatility distribution of isoprene oxidation products

61

under atmospheric conditions and thus on the importance of non-reactive gas-particle

62

partitioning of isoprene oxidation products as an SOA source.

63 64 65

1. Introduction Atmospheric submicron aerosol particles impact climate, human health, and visibility.1 A

66

large portion of submicron aerosol mass (~20-90%) is secondary organic aerosol (SOA),2,

67

which arises from chemical processes in the atmosphere that convert volatile organic compounds

68

(VOC) into condensed-phase organic material. A significant fraction of SOA is thought to be

69

formed from biogenic rather than anthropogenic VOC.4,

5

3

The annual emissions of isoprene

2 ACS Paragon Plus Environment

Page 3 of 30

Environmental Science & Technology

70

(C5H8), primarily from deciduous vegetation, represent nearly half the total of all non-methane

71

hydrocarbons.6 Thus, isoprene has the potential to generate large quantities of SOA, even if its

72

overall conversion is inefficient.

73

The formation of SOA from the free radical photochemical oxidation of biogenic VOC

74

(BVOC) is governed in part by the mass concentration (c) weighted volatility distribution of the

75

resulting products. For example, BVOC oxidation products having equilibrium saturation vapor

76

concentrations (c*) lower than ~100 µg m-3 can be expected to partition into an existing organic

77

condensed-phase and thus contribute to SOA, as typical organic aerosol mass concentrations

78

( ) are in the range of 1-10 µg m-3.7, 8 The volatility distribution of BVOC oxidation products

79

is governed by the ratio of the BVOC that undergoes functionalization relative to fragmentation

80

during atmospheric oxidation. Oxidation of BVOC is frequently initiated by the formation of a

81

carbon-centered radical primarily due to reaction with a hydroxyl radical (OH). In this initial

82

stage, atmospheric oxygen (O2) can add to the carbon-centered radical to form an organic peroxy

83

radical (RO2). RO2 can then undergo bimolecular reactions with typically HO2, NO, or RO2. All

84

of these bimolecular reactions can produce more functionalized products, while the latter two can

85

also form alkoxy radicals (RO) which can undergo H-shift isomerization9 or fragment into

86

smaller, more volatile products. Thus, the volatility distribution is largely determined by the fate

87

of RO2 intermediates.

88

In the specific case of isoprene (scheme 1), OH addition to a double bond followed by O2

89

addition yields a hydroxy RO2. Some isomers of the first-generation RO2 can undergo

90

unimolecular isomerization (not shown), with a forward rate constant of order 0.01 - 1 s-1,10-12

91

followed by HO2 loss to produce a hydroperoxyenal (HPALD),10, 11 or dihydroperoxy-carbonyl

92

peroxy radicals (di-HPCARP).12 The net isomerization rate constant weighted by the isomer

3 ACS Paragon Plus Environment

Environmental Science & Technology

12

Page 4 of 30

93

distribution was calculated to be ~0.001 s-1.11,

94

bimolecular reactions with HO2 to form a hydroxy hydroperoxide (ISOPOOH), NO to form an

95

alkyl nitrate,13 or NO or RO2 to form an alkoxy radical. The latter tends to decompose by carbon-

96

carbon bond scission to form products such as formaldehyde, methacrolein, or methyl vinyl

97

ketone.14-17 These first generation closed-shell products are rather volatile, and though they may

98

undergo multiphase reactions, their contribution to SOA by direct partitioning is likely small.

99

Given that a large fraction of isoprene reacts to form ISOPOOH under low NOx conditions,

100

possibly up to or greater than 70%,18 and that it likely has the lowest c* of the first generation

101

products, its oxidation products are the subject of continued focus.

The first generation RO2 can undergo

102

ISOPOOH reacts rapidly via OH addition to the remaining double bond producing a

103

carbon-centered radical often α to the hydroperoxide moiety (Scheme 1). This alkyl radical has

104

been shown to form an epoxide bridge with the hydroperoxide group. Consecutive loss of OH

105

from that hydroperoxide group forms an epoxydiol (IEPOX)18 ~70-80% of the time19 (Scheme

106

1). A significant fraction of isoprene-derived SOA is now understood to originate from acid

107

catalyzed aqueous phase chemistry of IEPOX,18, 20-27 although SOA also forms from isoprene

108

oxidation in the absence of an acidic aqueous phase.28-31 Recent gas-29 and particle-phase30-32

109

composition measurements of low-NOx OH oxidation of both isoprene and ISOPOOH reveal

110

highly oxygenated second generation products. A major particle-phase component has the

111

composition C5H12O6,30-32 likely a dihydroxy dihydroperoxide, which we refer to as

112

ISOP(OOH)2 (see Scheme 2, top reaction, for structure). This product is believed to form from

113

O2 addition to the carbon-centered radical from ISOPOOH + OH to yield a second generation

114

RO2 in competition with IEPOX formation (Scheme 1), followed by reaction of the RO2 with

115

HO2 (Scheme 2, top reaction). A fraction of ISOPOOH + OH is also expected to react via H-

4 ACS Paragon Plus Environment

Page 5 of 30

Environmental Science & Technology

116

abstraction from an alkyl group or the –OOH group, although these channels are relatively small

117

contributions to net ISOPOOH reactivity.19

118

In previous work, we showed that steady-state ISOP(OOH)2 and SOA mass yields

119

increase together as the OH and HO2 source,30 H2O2, is increased relative to isoprene. Reasons

120

for this behavior have yet to be determined. Mechanistic explanations for prominent products

121

from this system other than ISOP(OOH)2, such as C5H12O5 and C5H10O5-7, are also lacking.

122

Herein, we compare an expanded set of chamber experiments with novel predictions from a

123

detailed multiphase chemical model, and quantum chemical calculations of RO2 radical fates, to

124

improve our understanding of the isoprene oxidation mechanism and thus the formation of

125

isoprene SOA from non-IEPOX oxidation products. Specifically, we test different hypotheses for

126

the fate of the C5H11O6• peroxy radical so that we can understand the dependence of non-IEPOX

127

isoprene SOA formation on chamber oxidation conditions. We conclude with a discussion of the

128

atmospheric implications of our findings.

129 130

2. Methods

131

2.1 Laboratory Experiments

132

A suite of online instruments were employed to monitor the gas- and particle-phase chemical

133

composition. Ozone (Thermo Environmental Instruments model 49C), NO/NO2/NOx (Thermo

134

Environmental Instruments model 42C), and isoprene (Ionicon proton-transfer-reaction mass

135

spectrometer (PTR-MS)) concentrations were measured in the chamber outflow. Bulk submicron

136

organic and inorganic particle-phase composition and mass loading were measured with an

137

Aerodyne high-resolution time-of-flight aerosol mass spectrometer (HRToF-AMS). A high-

138

resolution time-of-flight chemical ionization mass spectrometer (HRToF-CIMS) with iodide

5 ACS Paragon Plus Environment

Environmental Science & Technology

Page 6 of 30

139

ionization was coupled to a Filter Inlet for Gases and AEROsols (FIGAERO).33 The HRToF-

140

CIMS FIGAERO coupling has been described in detail previously in the configuration used in

141

these experiments.30, 32, 34 Briefly, the FIGAERO is an inlet manifold that allows for sampling

142

gas-phase composition at 1 Hz through an independent inlet while collecting aerosol particles on

143

a Teflon filter through another inlet. Particle-phase composition is assessed at approximately

144

hourly resolution by actuating the filter to be in front of the HRToF-CIMS entrance orifice and

145

performing temperature-programmed thermal desorption of particles. Vaporized particle-phase

146

compounds are then detected by the HRToF-CIMS.

147

Experiments presented here have been discussed previously,32 but the observed

148

relationships of SOA yields to radical sources in these specific experiments have not been

149

presented. In brief, experiments were performed in 2015 in the Pacific Northwest National

150

Laboratory’s 10.6 m3 fluorinated ethylene polypropylene (FEP) environmental chamber, which

151

has been described before.32, 35 The chamber was controlled to a constant 50% relative humidity

152

and operated in continuous-flow mode with a residence time of 5.2 hours where reactants are

153

continuously delivered at a constant flow rate while products are continuously monitored with a

154

suite of gas- and particle-phase instrumentation. OH and HO2 radicals were generated from the

155

photolysis of excess H2O2 (Sigma-Aldrich, 50% in water), with chamber concentrations ranging

156

from 0.1 s-1.

, which is slower than other recent determinations in chambers operated in batch-mode.45-48, 50, 51

290

The important role of isomerization is demonstrated further in Figure 2, where we

291

compare model predicted and observed ISOP(OOH)2 mass yields as a function of H2O2 in the

292

chamber, used as a proxy for [HO2] (Fig. S2). Model runs were conducted at isomerization rate

293

constants, kisom, of 0 (i.e. the base-case), 0.005, 0.3, and 0.8 s-1 and all other parameters the same

294

as the base-case (IEPOX yield of 82%, kwall =10-5 s-1). The runs at the two slowest isomerization

295

rates (0 and 0.005 s-1) are statistically the same and significantly over-predict the ISOP(OOH)2

296

mass yield relative to the measurements by up to a factor of 30. The runs with kisom of 0.3 and 0.8

297

s-1 are both much closer to the data, with the run at kisom = 0.8 s-1 predicting ISOP(OOH)2 mass

298

yields closest to the observations (Fig. 2, inset), and have the approximate linear increase of yield

299

with H2O2. We demonstrate in the SI with a steady-state chemical analysis that the dependence

12 ACS Paragon Plus Environment

Page 13 of 30

Environmental Science & Technology

300

of the ISOP(OOH)2 and SOA mass yields on changing input H2O2 concentrations requires the

301

existence of a pathway such as isomerization that competes with the RO2 + HO2 reaction. Simply

302

increasing kwall or ϕ moves the model curve down, but does not change the shape (Fig. S4), nor

303

does [OH] change enough to explain the observations. We can also rule out significant nitric

304

oxide chemistry using our observations of organic nitrates (and NOx).32 Increasing the RO2 +

305

RO2 reaction rate constant to the kinetic limit also did not improve the agreement. Thus, we

306

conclude that isomerization of the C5H11O6• peroxy radical competes with the reaction of

307

C5H11O6• with HO2 and thereby causes the observed trend with H2O2 concentration.

308

While Crounse et al.11 reported experimentally measured net isomer-weighted

309

isomerization rate of the first generation isoprene hydroxy peroxy radical, C5H9O3 (ISOPO2), of

310

order 0.001 s-1, we note specific isomers had larger isomerization rates,73 and other studies have

311

found the H-shift kisom of methacrolein and a peroxy radical of -OOH substituted 3-pentanone to

312

be 0.5 s-1 74 and >0.1 s-1,73 respectively. Computational studies have also calculated rates similar

313

to these for peroxy radical hydrogen shift reactions.54, 75 Thus, the kisom required for agreement

314

with the observations we present is consistent with kisom for other peroxy radicals.

315

Our independently performed quantum chemical calculations examined the isomerization

316

rate of C5H11O6• and possible products for the two most prominent ISOPOOH isomers, 1,2-

317

ISOPOOH (shown in Scheme 1) and 3,4-ISOPOOH. In Scheme 3, we show the C5H11O6•

318

isomers formed from 1,2-ISOPOOH, however, the same is obtained for 3,4-ISOPOOH due to the

319

rapid interconversion between the two peroxy radicals, which has also been previously

320

demonstrated for another hydroperoxy-peroxy system.76 The forward rates of the 1,5 H-shift

321

reactions were found to be 0.3 s-1 and 0.9 s-1 for the two isomers, respectively (Scheme 3). The

322

reverse rates in both cases were too slow at room temperature to compete with epoxide formation

13 ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 30

323

or the addition of O2 to the resulting alkyl radical, which presumably occurs near the kinetic limit

324

(~108 s-1 at 1 atm total pressure) as previously shown.77 Thus, the ISOPOOH-derived RO2 will

325

convert into other products due to intramolecular H-shifts at a net forward rate of order 0.3 s-1 or

326

faster according to the quantum chemical calculations, consistent with the observations and box

327

modeling described above. The fact that the box model and quantum chemical calculations

328

independently arrive at a similar value for the rate of C5H11O6• isomerization supports the

329

hypothesis that such a pathway competes with HO2, not only in the chamber, but also the

330

atmosphere where HO2 concentrations are significantly lower.

331

For isomerization of the ISOPOOH-derived RO2 to explain a decreasing SOA mass yield,

332

the products that result from the isomerization must be of higher volatility than ISOP(OOH)2.

333

The quantum chemical calculations predict three possible products: a hydroxy dihydroperoxy

334

aldehyde or ketone (C5H10O6 or C5H10O7, respectively) and a dihydroxy hydroperoxy epoxide

335

(C5H10O5), (Schemes 2 and 3). Additionally, C5H10O5 could be a dihydroxy hydroperoxy

336

aldehyde formed from addition of OH to the internal carbon of the remaining ISOPOOH double

337

bond (Scheme S1). Previous studies suggest branching to this pathway is a few percent,13, 78, 79

338

and so was not included. As it is also a product of isomerization, it does not affect our

339

determination of kisom or conclusions (see SI). The calculations suggest that formation of the

340

epoxide, analogous mechanistically to the formation of IEPOX,18 is one to two orders of

341

magnitude faster than O2 addition to form the aldehyde/ketone, where O2 addition occurs at a

342

rate of ~107 s-1 for both 1,2- and 3,4-ISOPOOH and epoxide formation occurs at calculated rates

343

of 9 × 109 or 3 × 108 s-1 for 1,2- and 3,4-ISOPOOH, respectively. That an epoxide is formed is

344

intriguing, suggesting that epoxide formation may be a more general characteristic of isoprene

345

oxidation rather than specific to IEPOX, and that epoxide formation when a hydroperoxide is α

14 ACS Paragon Plus Environment

Page 15 of 30

Environmental Science & Technology

346

to a carbon-centered radical is more common than our current mechanisms of VOC oxidation

347

suggests.

348

Although the FIGAERO HRToF-CIMS does not provide structural information, we can

349

assess whether the products of C5H11O6• peroxy radical isomerization are consistent with the

350

observed compositions of measured oxidation products. Figure 3 shows the behavior of the sum

351

of isomerization products relative to the HO2 bimolecular reaction product, where we assume

352

ions with the compositions C5H10O5, C5H10O6, and C5H10O7 correspond to the predicted epoxide,

353

aldehyde, and ketone products, respectively. The model predicts the relative shape of the

354

relationship with respect to H2O2 correctly, but the absolute agreement in the ratio of these

355

products is better for kisom of ~0.1 s-1, as opposed to the 0.3 – 0.8 s-1 suggested by the comparison

356

to the concentration of ISOP(OOH)2. Uncertainties in the relative loss rates assumed in the

357

model and in the relative instrument response to these various products likely explain such

358

discrepancies. The results shown in Figures 2 and 3, combined with the quantum calculations, all

359

point to a kisom of 0.1-0.8 s-1. That independent approaches all arrive at kisom > 0.1 s-1 therefore

360

has important implications for SOA formation from isoprene in the atmosphere.

361

The SOA mass concentration and the distribution of compounds in the particle-phase

362

provide an additional test of the model. Figure 4, top, shows that a C5H11O6• kisom of 0.3 s-1

363

provides good model/measurement agreement of the SOA abundance (R2=0.98). The base-case

364

model over-predicts SOA, while kisom of 0.8 s-1 slightly under-predicts SOA. Consistent with the

365

observations, the modeled SOA contains 36% ISOP(OOH)2 (Fig. 4, bottom), with other major

366

components matching compositions of major components measured by the FIGAERO HRToF-

367

CIMS. The model predicts that the dihydroxy hydroperoxy epoxide formed from C5H11O6•

15 ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 30

368

isomerization makes up 48% of the aerosol composition at the highest HO2 concentration and

369

kisom of 0.8 s-1, while it was observed to be 11%.

370

We do not expect perfect agreement between measured and modeled SOA abundance and

371

composition, in part because previous studies have shown that up to 50% of this non-IEPOX

372

SOA is possibly low volatility accretion-like products32 likely formed from multiphase chemistry

373

which is not included in the model. Thus, we expect to under-predict measured SOA by as much

374

as a factor of two with F0AM, as well as to be unable to simulate observed products which are

375

the result of thermal decomposition of low volatility oligomers during the desorption analysis.

376

As such, we conclude that the overall agreement with observations in the predicted SOA mass

377

abundance, dominant products, and response to changing H2O2 is reasonable, and that

378

constraining kisom for the C5H11O6• peroxy radical at >0.1 s-1 is more important than kwall for

379

predicting the SOA yield.

380 381

4. Atmospheric Implications

382

We utilize quantum chemical calculations and a detailed chemical kinetics box model

383

compared to measurements of gas- and particle-phase oxidation products of isoprene made in an

384

environmental chamber. To explain the responses of product distributions and SOA mass

385

concentrations to changes in chamber radical conditions, the C5H11O6• peroxy radical formed

386

from ISOPOOH + OH must undergo intra-molecular H-atom shift reactions (“RO2

387

isomerization”) at rates exceeding 0.1 s-1 to form products 1-2 orders of magnitude higher

388

volatility than that of ISOP(OOH)2, which forms if the peroxy radical reacts with HO2. Such a

389

pathway has significant implications for how low-NO chamber studies should be conducted, and

16 ACS Paragon Plus Environment

Page 17 of 30

Environmental Science & Technology

390

offers a way to reconcile differences in SOA yields measured under different chamber

391

conditions.28-31

392

While ISOP(OOH)2 and related highly oxygenated C5 compounds believed to come from the

393

ISOP(OOH)2 pathway, have been observed in the Southeastern U.S. atmosphere during the

394

SOAS 2013 field campaign, their abundance was only a few percent of the total SOA in an

395

isoprene dominated region.31 Using SOAS diurnally averaged measurements of HO2 and NO,

396

and our F0AM model presented here, we calculate the vast majority, >96%, of the C5H11O6• will

397

undergo isomerization instead of forming the lowest volatility ISOP(OOH)2 and related products.

398

The quantum calculations suggest the most likely product of the isomerization is a dihydroxy

399

hydroperoxy epoxide. While we and others observe a major second generation product with the

400

composition (C5H10O5) of such an epoxide,29 confirmation of the structure remains necessary.

401

Epoxides in general undergo heterogeneous uptake and acid catalyzed ring opening with the

402

potential to form SOA through subsequent substitution reactions.27, 80, 81 However, we speculate

403

that the dominant fate in aqueous acidic particles of this specific epoxide moiety will be nearly

404

instantaneous conversion to a dihydroxy hydroperoxy carbonyl, but direct experimental evidence

405

is needed. That said, its lower volatility relative to IEPOX will allow it to partition more strongly

406

to an organic phase where it may undergo accretion chemistry.

407

The dominance of epoxide formation over O2 addition to a carbon radical α to a

408

hydroperoxyl group, as calculated herein and for the case of IEPOX, suggests gas-phase epoxide

409

formation may be more common than current atmospheric chemistry mechanisms allow. Such a

410

situation would have implications for the propensity of acidity enhanced SOA formation from a

411

variety of biogenic and anthropogenic VOC.20,

412

autoxidation, where peroxy radicals form multiple hydroperoxide moieties by several intra-

26, 27, 81

For example, the peroxy radical

17 ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 30

413

molecular H-shifts could well terminate to epoxides. Epoxide formation has been commonly

414

suggested, particularly for aromatic systems,82-88 supporting the idea that epoxides are potentially

415

a feature of low NOx oxidation of non-aromatics when a radical center occurs α to a hydroperoxy

416

group. Moreover, there is also recent evidence for epoxide formation from –ONO2 groups α to

417

the carbon radical.19, 89, 90 Thus, we suggest analytical methods specific to epoxides be developed

418

for assessing their ubiquity in the atmosphere and chamber studies.

419 420

Associated Content

421

Includes further detail on: chamber conditions, F0AM and the gas/particle partitioning module

422

used herein, quantum chemical calculations, and the dependence of ISOP(OOH)2 and SOA mass

423

yields on H2O2. Also includes a link to the gas/particle partitioning module for download.

424 425

Acknowledgements

426

This work was supported by the U.S. Department of Energy ASR grants DE-SC0011791.

427

E.L.D. was supported by the National Science Foundation Graduate Research Fellowship under

428

Grant No. DGE-1256082. B.H.L. was supported by the National Oceanic and Atmospheric

429

Administration (NOAA) Climate and Global Change Postdoctoral Fellowship Program. PNNL

430

authors were supported by the U. S. Department of Energy, Office of Biological and

431

Environmental Research, as part of the ASR program. Copenhagen authors were supported by

432

the University of Copenhagen. Pacific Northwest National Laboratory is operated for the DOE

433

by Battelle Memorial Institute under contract DE-AC05-76RL01830. We thank G.M. Wolfe

434

(NASA/UMBC), T.A. Spencer (Dartmouth College), R. Ditchfield (Dartmouth College), and

18 ACS Paragon Plus Environment

Page 19 of 30

Environmental Science & Technology

435

Rasmus V. Otkjær (University of Copenhagen) for helpful discussions, and the Danish Center

436

for Scientific Computing for computational time.

437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 19 ACS Paragon Plus Environment

Environmental Science & Technology

479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526

Page 20 of 30

References 1. Poschl, U., Atmospheric aerosols: Composition, transformation, climate and health effects. Angew. Chem.-Int. Edit. 2005, 44 (46), 7520-7540; doi 10.1002/anie.200501122. 2. Jimenez, J. L.; Canagaratna, M. R.; Donahue, N. M.; Prevot, A. S. H.; Zhang, Q.; Kroll, J. H.; DeCarlo, P. F.; Allan, J. D.; Coe, H.; Ng, N. L.; Aiken, A. C.; Docherty, K. S.; Ulbrich, I. M.; Grieshop, A. P.; Robinson, A. L.; Duplissy, J.; Smith, J. D.; Wilson, K. R.; Lanz, V. A.; Hueglin, C.; Sun, Y. L.; Tian, J.; Laaksonen, A.; Raatikainen, T.; Rautiainen, J.; Vaattovaara, P.; Ehn, M.; Kulmala, M.; Tomlinson, J. M.; Collins, D. R.; Cubison, M. J.; Dunlea, E. J.; Huffman, J. A.; Onasch, T. B.; Alfarra, M. R.; Williams, P. I.; Bower, K.; Kondo, Y.; Schneider, J.; Drewnick, F.; Borrmann, S.; Weimer, S.; Demerjian, K.; Salcedo, D.; Cottrell, L.; Griffin, R.; Takami, A.; Miyoshi, T.; Hatakeyama, S.; Shimono, A.; Sun, J. Y.; Zhang, Y. M.; Dzepina, K.; Kimmel, J. R.; Sueper, D.; Jayne, J. T.; Herndon, S. C.; Trimborn, A. M.; Williams, L. R.; Wood, E. C.; Middlebrook, A. M.; Kolb, C. E.; Baltensperger, U.; Worsnop, D. R., Evolution of organic aerosols in the atmosphere. Science 2009, 326 (5959), 1525-1529; doi 10.1126/science.1180353. 3. Zhang, Q.; Jimenez, J. L.; Canagaratna, M. R.; Allan, J. D.; Coe, H.; Ulbrich, I.; Alfarra, M. R.; Takami, A.; Middlebrook, A. M.; Sun, Y. L.; Dzepina, K.; Dunlea, E.; Docherty, K.; DeCarlo, P. F.; Salcedo, D.; Onasch, T.; Jayne, J. T.; Miyoshi, T.; Shimono, A.; Hatakeyama, S.; Takegawa, N.; Kondo, Y.; Schneider, J.; Drewnick, F.; Borrmann, S.; Weimer, S.; Demerjian, K.; Williams, P.; Bower, K.; Bahreini, R.; Cottrell, L.; Griffin, R. J.; Rautiainen, J.; Sun, J. Y.; Zhang, Y. M.; Worsnop, D. R., Ubiquity and dominance of oxygenated species in organic aerosols in anthropogenically-influenced Northern Hemisphere midlatitudes. Geophys. Res. Lett. 2007, 34 (13), L13801; doi 10.1029/2007gl029979. 4. Weber, R. J.; Sullivan, A. P.; Peltier, R. E.; Russell, A.; Yan, B.; Zheng, M.; de Gouw, J.; Warneke, C.; Brock, C.; Holloway, J. S.; Atlas, E. L.; Edgerton, E., A study of secondary organic aerosol formation in the anthropogenic-influenced southeastern United States. J. Geophys. Res.-Atmos. 2007, 112 (D13), D13302; doi 10.1029/2007jd008408. 5. Schichtel, B. A.; Malm, W. C.; Bench, G.; Fallon, S.; McDade, C. E.; Chow, J. C.; Watson, J. G., Fossil and contemporary fine particulate carbon fractions at 12 rural and urban sites in the United States. J. Geophys. Res.-Atmos. 2008, 113 (D2), D02311; doi 10.1029/2007jd008605. 6. Guenther, A. B.; Jiang, X.; Heald, C. L.; Sakulyanontvittaya, T.; Duhl, T.; Emmons, L. K.; Wang, X., The model of emissions of gases and aerosols from nature version 2.1 (MEGAN2.1): an extended and updated framework for modeling biogenic emissions. Geosci. Model Dev. 2012, 5 (6), 1471-1492; doi 10.5194/gmd-5-1471-2012. 7. Pankow, J. F., An absorption-model of the gas aerosol partitioning involved in the formation of secondary organic aerosol Atmos. Environ. 1994, 28 (2), 189-193; doi 10.1016/1352-2310(94)90094-9. 8. Donahue, N. M.; Trump, E. R.; Pierce, J. R.; Riipinen, I., Theoretical constraints on pure vaporpressure driven condensation of organics to ultrafine particles. Geophys. Res. Lett. 2011, 38 (16), L16801; doi 10.1029/2011gl048115. 9. Atkinson, R.; Lloyd, A. C., Evaluation of kinetic and mechanistic data for modeling of photochemical smog. J. Phys. Chem. Ref. Data 1984, 13 (2), 315-444. 10. Peeters, J.; Nguyen, T. L.; Vereecken, L., HOx radical regeneration in the oxidation of isoprene. Phys. Chem. Chem. Phys. 2009, 11 (28), 5935-5939; doi 10.1039/b908511d. 11. Crounse, J. D.; Paulot, F.; Kjaergaard, H. G.; Wennberg, P. O., Peroxy radical isomerization in the oxidation of isoprene. Phys. Chem. Chem. Phys. 2011, 13 (30), 13607-13613; doi 10.1039/c1cp21330j. 12. Peeters, J.; Muller, J. F.; Stavrakou, T.; Nguyen, V. S., Hydroxyl radical recycling in isoprene oxidation driven by hydrogen bonding and hydrogen tunneling: the upgraded LIM1 mechanism. J. Phys. Chem. A 2014, 118 (38), 8625-43; doi 10.1021/jp5033146. 13. Paulot, F.; Crounse, J. D.; Kjaergaard, H. G.; Kroll, J. H.; Seinfeld, J. H.; Wennberg, P. O., Isoprene photooxidation: new insights into the production of acids and organic nitrates. Atmos. Chem. Phys. 2009, 9 (4), 1479-1501. 20 ACS Paragon Plus Environment

Page 21 of 30

527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573

Environmental Science & Technology

14. Miyoshi, A.; Hatakeyama, S.; Washida, N., OH radical-initiated photooxidation of isoprene- an estimate of global CO production. J. Geophys. Res.-Atmos. 1994, 99 (D9), 18779-18787; doi 10.1029/94jd01334. 15. Paulson, S. E.; Seinfeld, J. H., Development and evaluation of a photooxidation mechanism for isoprene J. Geophys. Res.-Atmos. 1992, 97 (D18), 20703-20715. 16. Atkinson, R.; Aschmann, S. M.; Tuazon, E. C.; Arey, J.; Zielinska, B., Formation of 3-methylfuran from the gas-phase reaction of OH radicals with isoprene and the rate-constant for its reaction with the OH radical Int. J. Chem. Kinet. 1989, 21 (7), 593-604; doi 10.1002/kin.550210709. 17. Tuazon, E. C.; Atkinson, R., A product study of the gas-phase reaction of isoprene with the OH radical in the presence of NOx. Int. J. Chem. Kinet. 1990, 22 (12), 1221-1236; doi 10.1002/kin.550221202. 18. Paulot, F.; Crounse, J. D.; Kjaergaard, H. G.; Kurten, A.; St Clair, J. M.; Seinfeld, J. H.; Wennberg, P. O., Unexpected epoxide formation in the gas-phase photooxidation of isoprene. Science 2009, 325 (5941), 730-733; doi 10.1126/science.1172910. 19. St Clair, J. M.; Rivera-Rios, J. C.; Crounse, J. D.; Knap, H. C.; Bates, K. H.; Teng, A. P.; Jørgensen, S.; Kjaergaard, H. G.; Keutsch, F. N.; Wennberg, P. O., Kinetics and products of the reaction of the firstgeneration isoprene hydroxy hydroperoxide (ISOPOOH) with OH. J. Phys. Chem. A 2016, 120 (9), 14411451; doi 10.1021/acs.jpca.5b06532. 20. Gaston, C. J.; Riedel, T. P.; Zhang, Z. F.; Gold, A.; Surratt, J. D.; Thornton, J. A., Reactive uptake of an isoprene-derived epoxydiol to submicron aerosol particles. Environ. Sci. Technol. 2014, 48 (19), 11178-11186; doi 10.1021/es5034266. 21. Lin, Y. H.; Budisulistiorini, H.; Chu, K.; Siejack, R. A.; Zhang, H. F.; Riva, M.; Zhang, Z. F.; Gold, A.; Kautzman, K. E.; Surratt, J. D., Light-absorbing oligomer formation in secondary organic aerosol from reactive uptake of isoprene epoxydiols. Environ. Sci. Technol. 2014, 48 (20), 12012-12021; doi 10.1021/es503142b. 22. Lin, Y. H.; Knipping, E. M.; Edgerton, E. S.; Shaw, S. L.; Surratt, J. D., Investigating the influences of SO2 and NH3 levels on isoprene-derived secondary organic aerosol formation using conditional sampling approaches. Atmos. Chem. Phys. 2013, 13 (16), 8457-8470; doi 10.5194/acp-13-8457-2013. 23. Lin, Y. H.; Zhang, H. F.; Pye, H. O. T.; Zhang, Z. F.; Marth, W. J.; Park, S.; Arashiro, M.; Cui, T. Q.; Budisulistiorini, H.; Sexton, K. G.; Vizuete, W.; Xie, Y.; Luecken, D. J.; Piletic, I. R.; Edney, E. O.; Bartolotti, L. J.; Gold, A.; Surratt, J. D., Epoxide as a precursor to secondary organic aerosol formation from isoprene photooxidation in the presence of nitrogen oxides. Proc. Natl. Acad. Sci. U. S. A. 2013, 110 (17), 67186723; doi 10.1073/pnas.1221150110. 24. Lin, Y. H.; Zhang, Z. F.; Docherty, K. S.; Zhang, H. F.; Budisulistiorini, S. H.; Rubitschun, C. L.; Shaw, S. L.; Knipping, E. M.; Edgerton, E. S.; Kleindienst, T. E.; Gold, A.; Surratt, J. D., Isoprene epoxydiols as precursors to secondary organic aerosol formation: Acid-catalyzed reactive uptake studies with authentic compounds. Environ. Sci. Technol. 2012, 46 (1), 250-258; doi 10.1021/es202554c. 25. Liu, Y.; Kuwata, M.; Strick, B. F.; Geiger, F. M.; Thomson, R. J.; McKinney, K. A.; Martin, S. T., Uptake of epoxydiol isomers accounts for half of the particle-phase material produced from isoprene photooxidation via the HO pathway. Environ. Sci. Technol. 2014, 49, 250-258; doi 10.1021/es5034298. 26. Surratt, J. D.; Chan, A. W. H.; Eddingsaas, N. C.; Chan, M. N.; Loza, C. L.; Kwan, A. J.; Hersey, S. P.; Flagan, R. C.; Wennberg, P. O.; Seinfeld, J. H., Reactive intermediates revealed in secondary organic aerosol formation from isoprene. Proc. Natl. Acad. Sci. U. S. A. 2010, 107 (15), 6640-6645; doi 10.1073/pnas.0911114107. 27. Surratt, J. D.; Murphy, S. M.; Kroll, J. H.; Ng, N. L.; Hildebrandt, L.; Sorooshian, A.; Szmigielski, R.; Vermeylen, R.; Maenhaut, W.; Claeys, M.; Flagan, R. C.; Seinfeld, J. H., Chemical composition of secondary organic aerosol formed from the photooxidation of isoprene. J. Phys. Chem. A 2006, 110 (31), 9665-9690; doi 10.1021/jp061734m.

21 ACS Paragon Plus Environment

Environmental Science & Technology

574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621

Page 22 of 30

28. Kroll, J. H.; Ng, N. L.; Murphy, S. M.; Flagan, R. C.; Seinfeld, J. H., Secondary organic aerosol formation from isoprene photooxidation. Environ. Sci. Technol. 2006, 40 (6), 1869-1877; doi 10.1021/es0524301. 29. Krechmer, J. E.; Coggon, M. M.; Massoli, P.; Nguyen, T. B.; Crounse, J. D.; Hu, W. W.; Day, D. A.; Tyndall, G. S.; Henze, D. K.; Rivera-Rios, J. C.; Nowak, J. B.; Kimmel, J. R.; Mauldin, R. L.; Stark, H.; Jayne, J. T.; Sipila, M.; Junninen, H.; Clair, J. M. S.; Zhang, X.; Feiner, P. A.; Zhang, L.; Miller, D. O.; Brune, W. H.; Keutsch, F. N.; Wennberg, P. O.; Seinfeld, J. H.; Worsnop, D. R.; Jimenez, J. L.; Canagaratna, M. R., Formation of low volatility organic compounds and Secondary Organic Aerosol from isoprene hydroxyhydroperoxide low-NO oxidation. Environ. Sci. Technol. 2015, 49 (17), 10330-10339; doi 10.1021/acs.est.5b02031. 30. Liu, J. M.; D'Ambro, E. L.; Lee, B. H.; Lopez-Hilfiker, F. D.; Zaveri, R. A.; Rivera-Rios, J. C.; Keutsch, F. N.; Iyer, S.; Kurten, T.; Zhang, Z. F.; Gold, A.; Surratt, J. D.; Shilling, J. E.; Thornton, J. A., Efficient isoprene secondary organic aerosol formation from a non-IEPOX pathway. Environ. Sci. Technol. 2016, 50 (18), 9872-9880; doi 10.1021/acs.est.6b01872. 31. Riva, M.; Budisulistiorini, S. H.; Chen, Y. Z.; Zhang, Z. F.; D'Ambro, E. L.; Zhang, X.; Gold, A.; Turpin, B. J.; Thornton, J. A.; Canagaratna, M. R.; Surratt, J. D., Chemical characterization of secondary organic aerosol from oxidation of isoprene hydroxyhydroperoxides. Environ. Sci. Technol. 2016, 50 (18), 9889-9899; doi 10.1021/acs.est.6b02511. 32. D'Ambro, E. L.; Lee, B. H.; Liu, J.; Shilling, J. E.; Gaston, C. J.; Lopez-Hilfiker, F. D.; Schobesberger, S.; Zaveri, R. A.; Mohr, C.; Lutz, A.; Zhang, Z.; Gold, A.; Surratt, J. D.; Rivera-Rios, J. C.; Keutsch, F. N.; Thornton, J. A., Molecular composition and volatility of isoprene photochemical oxidation secondary organic aerosol under low- and high-NOx conditions. Atmos. Chem. Phys. 2017, 17 (1), 159-174; doi 10.5194/acp-17-159-2017. 33. Lopez-Hilfiker, F. D.; Mohr, C.; Ehn, M.; Rubach, F.; Kleist, E.; Wildt, J.; Mentel, T. F.; Lutz, A.; Hallquist, M.; Worsnop, D.; Thornton, J. A., A novel method for online analysis of gas and particle composition: description and evaluation of a Filter Inlet for Gases and AEROsols (FIGAERO). Atmos. Meas. Tech. 2014, 7 (4), 983-1001; doi 10.5194/amt-7-983-2014. 34. Lee, B. H.; Lopez-Hilfiker, F. D.; Mohr, C.; Kurten, T.; Worsnop, D. R.; Thornton, J. A., An iodideadduct high-resolution time-of-flight chemical-ionization mass spectrometer: application to atmospheric inorganic and organic compounds. Environ. Sci. Technol. 2014, 48 (11), 6309-17; doi 10.1021/es500362a. 35. Liu, S.; Shilling, J. E.; Song, C.; Hiranuma, N.; Zaveri, R. A.; Russell, L. M., Hydrolysis of organonitrate functional groups in aerosol particles. Aerosol Sci. Technol. 2012, 46 (12), 1359-1369; doi 10.1080/02786826.2012.716175. 36. Wolfe, G. M.; Marvin, M. R.; Roberts, S. J.; Travis, K. R.; Liao, J., The Framework for 0-D Atmospheric Modeling (F0AM) v3.1. Geosci. Model Dev. 2016, 9 (9), 3309-3319; doi 10.5194/gmd-93309-2016. 37. Wolfe, G. M.; Thornton, J. A., The Chemistry of Atmosphere-Forest Exchange (CAFE) Model Part 1: Model description and characterization. Atmos. Chem. Phys. 2011, 11 (1), 77-101; doi 10.5194/acp-11-77-2011. 38. Wolfe, G. M.; Thornton, J. A.; Bouvier-Brown, N. C.; Goldstein, A. H.; Park, J. H.; McKay, M.; Matross, D. M.; Mao, J.; Brune, W. H.; LaFranchi, B. W.; Browne, E. C.; Min, K. E.; Wooldridge, P. J.; Cohen, R. C.; Crounse, J. D.; Faloona, I. C.; Gilman, J. B.; Kuster, W. C.; de Gouw, J. A.; Huisman, A.; Keutsch, F. N., The Chemistry of Atmosphere-Forest Exchange (CAFE) Model - Part 2: Application to BEARPEX-2007 observations. Atmos. Chem. Phys. 2011, 11 (3), 1269-1294; doi 10.5194/acp-11-12692011. 39. Wolfe, G. M.; Thornton, J. A.; McKay, M.; Goldstein, A. H., Forest-atmosphere exchange of ozone: sensitivity to very reactive biogenic VOC emissions and implications for in-canopy photochemistry. Atmos. Chem. Phys. 2011, 11 (15), 7875-7891; doi 10.5194/acp-11-7875-2011. 22 ACS Paragon Plus Environment

Page 23 of 30

622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667

Environmental Science & Technology

40. Jenkin, M. E.; Young, J. C.; Rickard, A. R., The MCM v3.3.1 degradation scheme for isoprene. Atmos. Chem. Phys. 2015, 15 (20), 11433-11459; doi 10.5194/acp-15-11433-2015. 41. Hilal, S. H.; Karickhoff, S. W.; Carreira, L. A., Prediction of the vapor pressure boiling point, heat of vaporization and diffusion coefficient of organic compounds. QSAR Comb. Sci. 2003, 22 (6), 565-574; doi 10.1002/qsar.200330812. 42. Saleh, R.; Donahue, N. M.; Robinson, A. L., Time scales for gas-particle partitioning equilibration of secondary organic aerosol formed from alpha-pinene ozonolysis. Environ. Sci. Technol. 2013, 47 (11), 5588-5594; doi 10.1021/es400078d. 43. Compernolle, S.; Ceulemans, K.; Muller, J. F., EVAPORATION: a new vapour pressure estimation method for organic molecules including non-additivity and intramolecular interactions. Atmos. Chem. Phys. 2011, 11 (18), 9431-9450; doi 10.5194/acp-11-9431-2011. 44. Crump, J. G.; Flagan, R. C.; Seinfeld, J. H., Particle wall loss rates in vessels. Aerosol Sci. Technol. 1983, 2 (3), 303-309; doi 10.1080/02786828308958636. 45. 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; doi 10.1073/pnas.1404727111. 46. 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; doi 10.1021/acs.est.6b00606. 47. 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; doi 10.5194/acp-16-1417-2016. 48. 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; doi 10.1080/02786826.2010.501044. 49. Loza, C. L.; Chan, A. W. H.; Galloway, M. M.; Keutsch, F. N.; Flagan, R. C.; Seinfeld, J. H., Characterization of vapor wall loss in laboratory chambers. Environ. Sci. Technol. 2010, 44 (13), 50745078; doi 10.1021/es100727v. 50. Yeh, G. K.; Ziemann, P. J., Alkyl nitrate formation from the reactions of C-8-C-14 n-alkanes with OH radicals in the presence of NOx: Measured yields with essential corrections for gas-wall partitioning. J. Phys. Chem. A 2014, 118 (37), 8147-8157; doi 10.1021/jp500631v. 51. McMurry, P. H.; Grosjean, D., Gas and aerosol wall losses in teflon film smog chambers Environ. Sci. Technol. 1985, 19 (12), 1176-1182; doi 10.1021/es00142a006. 52. Shilling, J. E.; Chen, Q.; King, S. M.; Rosenoern, T.; Kroll, J. H.; Worsnop, D. R.; McKinney, K. A.; Martin, S. T., Particle mass yield in secondary organic aerosol formed by the dark ozonolysis of alphapinene. Atmos. Chem. Phys. 2008, 8 (7), 2073-2088; doi 10.5194/acp-8-2073-2008. 53. Vereecken, L.; Peeters, J., The 1,5-H-shift in 1-butoxy: A case study in the rigorous implementation of transition state theory for a multirotamer system. J. Chem. Phys. 2003, 119 (10), 5159-5170; doi 10.1063/1.1597479. 54. Møller, K. H.; Otkjær, R. V.; Hyttinen, N.; Kurtén, T.; Kjaergaard, H. G., Cost-Effective Implementation of Multiconformer Transition State Theory for Peroxy Radical Hydrogen Shift Reactions. J. Phys. Chem. A 2016, 120 (51), 10072-10087; doi 10.1021/acs.jpca.6b09370. 55. Spartan '14, Wavefunction Inc.: Irvine, CA, 2014. 56. Clark, M.; Cramer, R. D.; Vanopdenbosch, N., Validation of the general-purpose TRIPOS 5.2 force-field. J. Comput. Chem. 1989, 10 (8), 982-1012; doi 10.1002/jcc.540100804.

23 ACS Paragon Plus Environment

Environmental Science & Technology

668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715

Page 24 of 30

57. Clark, T.; Chandrasekhar, J.; Spitznagel, G. W.; Schleyer, P. V., Efficient diffuse functionaugmented basis sets for anion calculations. III. The 3-21+G basis set for first-row elements, Li-F J. Comput. Chem. 1983, 4 (3), 294-301; doi 10.1002/jcc.540040303. 58. Becke, A. D., Density-functional thermochemistry .3. The role of exact exchange. J. Chem. Phys. 1993, 98 (7), 5648-5652; doi 10.1063/1.464913. 59. Frisch, M. J.; Pople, J. A.; Binkley, J. S., Self-consistent molecular-orbital methods .25. Supplementary functions for gaussian-basis sets J. Chem. Phys. 1984, 80 (7), 3265-3269; doi 10.1063/1.447079. 60. Hehre, W. J.; Ditchfield, R.; Pople, J. A., Self-consistent molecular-orbital methods .12. Further extensions of gaussian-type basis sets for use in molecular-orbital studies of organic-molecules J. Chem. Phys. 1972, 56 (5), 2257-+; doi 10.1063/1.1677527. 61. Lee, C. T.; Yang, W. T.; Parr, R. G., Development of the Colle-Salvetti correlation-energy formula into a functional of the electron-density Phys. Rev. B 1988, 37 (2), 785-789; doi 10.1103/PhysRevB.37.785. 62. Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J., Gaussian 09 Revision D.01. 2009. 63. Chai, J.-D.; Head-Gordon, M., Long-range corrected hybrid density functionals with damped atom-atom dispersion corrections. Phys. Chem. Chem. Phys. 2008, 10 (44), 6615-6620; doi 10.1039/b810189b. 64. Dunning, T. H., Gaussian-basis sets for use in correlated molecular calculations .1. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90 (2), 1007-1023; doi 10.1063/1.456153. 65. Kendall, R. A.; Dunning, T. H.; Harrison, R. J., Electron-affinities of the 1st-row atoms revistedSystematic basis-sets and wave-functions J. Chem. Phys. 1992, 96 (9), 6796-6806; doi 10.1063/1.462569. 66. Adler, T. B.; Knizia, G.; Werner, H. J., A simple and efficient CCSD(T)-F12 approximation. J. Chem. Phys. 2007, 127 (22), doi 10.1063/1.2817618. 67. Knizia, G.; Adler, T. B.; Werner, H. J., Simplified CCSD(T)-F12 methods: Theory and benchmarks. J. Chem. Phys. 2009, 130 (5), doi 10.1063/1.3054300. 68. Peterson, K. A.; Adler, T. B.; Werner, H. J., Systematically convergent basis sets for explicitly correlated wavefunctions: The atoms H, He, B-Ne, and Al-Ar. J. Chem. Phys. 2008, 128 (8), doi 10.1063/1.2831537. 69. Watts, J. D.; Gauss, J.; Bartlett, R. J., Coupled-cluster methods with noniterative triple excitations for restricted open-shell Hartree-Fock and other general single determinant reference functions Energies and analytical gradients J. Chem. Phys. 1993, 98 (11), 8718-8733; doi 10.1063/1.464480. 70. Werner, H. J.; Knizia, G.; Manby, F. R., Explicitly correlated coupled cluster methods with pairspecific geminals. Mol. Phys. 2011, 109 (3), 407-417; doi 10.1080/00268976.2010.526641. 71. Werner, H. J.; Knowles, P. J.; Knizia, G.; Manby, F. R.; Schütz, M.; Celani, P.; Györffy, W.; Kats, D.; Korona, T.; Lindh, R.; Mitrushenkov, A.; Rauhut, G.; Shamasundar, K. R.; Adler, T. B.; Amos, R. D.; Bernhardsson, A.; Berning, A.; Cooper, D. L.; Deegan, M. J. O.; Dobbyn, A. J.; Eckert, F.; Goll, E.; Hampel, C.; Hesselmann, A.; Hetzer, G.; Hrenar, T.; Jansen, G.; Köppl, C.; Liu, Y.; Lloyd, A. W.; Mata, R. A.; May, A. J.; McNicholas, S. J.; Meyer, W.; Mura, M. E.; Nicklass, A.; O’Neill, D. P.; Palmieri, P.; Peng, D.; Pflüger, K.; 24 ACS Paragon Plus Environment

Page 25 of 30

716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763

Environmental Science & Technology

Pitzer, R.; Reiher, M.; Shiozaki, T.; Stoll, H.; Stone, A. J.; Tarroni, R.; Thorsteinsson, T.; Wang, M., Molpro, version 2012.1, a package of ab initio programs. 2012. 72. Eckart, C., The penetration of a potential barrier by electrons. Phys. Rev. 1930, 35 (11), 13031309; doi 10.1103/PhysRev.35.1303. 73. Crounse, J. D.; Nielsen, L. B.; Jørgensen, S.; Kjaergaard, H. G.; Wennberg, P. O., Autoxidation of organic compounds in the atmosphere. J. Phys. Chem. Lett. 2013, 4 (20), 3513-3520; doi 10.1021/jz4019207. 74. Crounse, J. D.; Knap, H. C.; Ornso, K. B.; Jørgensen, S.; Paulot, F.; Kjaergaard, H. G.; Wennberg, P. O., Atmospheric fate of methacrolein. 1. Peroxy radical isomerization following addition of OH and O-2. J. Phys. Chem. A 2012, 116 (24), 5756-5762; doi 10.1021/jp211560u. 75. Rissanen, M. P.; Kurten, T.; Sipila, M.; Thornton, J. A.; Kangasluoma, J.; Sarnela, N.; Junninen, H.; Jørgensen, S.; Schallhart, S.; Kajos, M. K.; Taipale, R.; Springer, M.; Mentel, T. F.; Ruuskanen, T.; Petaja, T.; Worsnop, D. R.; Kjaergaard, H. G.; Ehn, M., The formation of highly oxidized multifunctional products in the ozonolysis of cyclohexene. J. Am. Chem. Soc. 2014, 136 (44), 15596-15606; doi 10.1021/ja507146s. 76. Jørgensen, S.; Knap, H. C.; Otkjaer, R. V.; Jensen, A. M.; Kjeldsen, M. L. H.; Wennberg, P. O.; Kjaergaard, H. G., Rapid hydrogen shift scrambling in hydroperoxy-substituted organic peroxy radicals. J. Phys. Chem. A 2016, 120 (2), 266-275; doi 10.1021/acs.jpca.5b067613. 77. Park, J.; Jongsma, C. G.; Zhang, R. Y.; North, S. W., OH/OD initiated oxidation of isoprene in the presence of O-2 and NO. J. Phys. Chem. A 2004, 108 (48), 10688-10697; doi 10.1021/jp040421t. 78. Ziemann, P. J.; Atkinson, R., Kinetics, products, and mechanisms of secondary organic aerosol formation. Chem. Soc. Rev. 2012, 41 (19), 6582-6605; doi 10.1039/c2cs35122f. 79. Peeters, J.; Boullart, W.; Pultau, V.; Vandenberk, S.; Vereecken, L., Structure-activity relationship for the addition of OH to (poly)alkenes: Site-specific and total rate constants. J. Phys. Chem. A 2007, 111 (9), 1618-1631; doi 10.1021/jp066973o. 80. Surratt, J. D.; Lewandowski, M.; Offenberg, J. H.; Jaoui, M.; Kleindienst, T. E.; Edney, E. O.; Seinfeld, J. H., Effect of acidity on secondary organic aerosol formation from isoprene. Environ. Sci. Technol. 2007, 41 (15), 5363-5369; doi 10.1021/es0704176. 81. Eddingsaas, N. C.; VanderVelde, D. G.; Wennberg, P. O., Kinetics and products of the acidcatalyzed ring-opening of atmospherically relevant butyl epoxy alcohols. J. Phys. Chem. A 2010, 114 (31), 8106-8113; doi 10.1021/jp103907c. 82. Richters, S.; Herrmann, H.; Berndt, T., Different pathways of the formation of highly oxidized multifunctional organic compounds (HOMs) from the gas-phase ozonolysis of beta-caryophyllene. Atmos. Chem. Phys. 2016, 16 (15), 9831-9845; doi 10.5194/acp-16-9831-2016. 83. Yu, J. Z.; Jeffries, H. E., Atmospheric photooxidation of alkylbenzenes .2. Evidence of formation of epoxide intermediates. Atmos. Environ. 1997, 31 (15), 2281-2287; doi 10.1016/s1352-2310(97)886372. 84. Glowacki, D. R.; Wang, L. M.; Pilling, M. J., Evidence of formation of bicyclic species in the early stages of atmospheric benzene oxidation. J. Phys. Chem. A 2009, 113 (18), 5385-5396; doi 10.1021/jp9001466. 85. Bartolotti, L. J.; Edney, E. O., Density-functional theory derived intermediates from the OH initiated atmospheric oxidation of toluene. Chem. Phys. Lett. 1995, 245 (1), 119-122; doi 10.1016/00092614(95)00953-2. 86. Motta, F.; Ghigo, G.; Tonachini, G., Oxidative degradation of benzene in the troposphere. Theoretical mechanistic study of the formation of unsaturated dialdehydes and dialdehyde epoxides. J. Phys. Chem. A 2002, 106 (17), 4411-4422; doi 10.1021/jp015619h. 87. Suh, I.; Zhang, R. Y.; Molina, L. T.; Molina, M. J., Oxidation mechanism of aromatic peroxy and bicyclic radicals from OH-toluene reactions. J. Am. Chem. Soc. 2003, 125 (41), 12655-12665; doi 10.1021/ja0350280. 25 ACS Paragon Plus Environment

Environmental Science & Technology

764 765 766 767 768 769 770 771

Page 26 of 30

88. Pan, S. S.; Wang, L. M., Atmospheric oxidation mechanism of m-xylene initiated by OH radical. J. Phys. Chem. A 2014, 118 (45), 10778-10787; doi 10.1021/jp506815v. 89. Jacobs, M. I.; Burke, W. J.; Elrod, M. J., Kinetics of the reactions of isoprene-derived hydroxynitrates: gas phase epoxide formation and solution phase hydrolysis. Atmos. Chem. Phys. 2014, 14 (17), 8933-8946; doi 10.5194/acp-14-8933-2014. 90. Lee, L.; Teng, A. P.; Wennberg, P. O.; Crounse, J. D.; Cohen, R. C., On rates and mechanisms of OH and O3 reactions with isoprene-derived hydroxy nitrates. J. Phys. Chem. A 2014, 118 (9), 1622-37; doi 10.1021/jp4107603.

772 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792 793 794 795 796 797 798 799 800 801 802

Figures

26 ACS Paragon Plus Environment

Page 27 of 30

803 804 805 806

807

Environmental Science & Technology

Scheme 1. Isoprene photochemical oxidation mechanism for one isomer of ISOPOOH under low-NOx conditions. ϕ and 1-ϕ represent the branching ratio between IEPOX and the peroxy radical, respectively.

Scheme 2. Proposed mechanism for one isomer of C5H11O6, the peroxy radical formed from ISOPOOH + OH in scheme 1. The formation of the three isomerization products is shown in detail in Scheme 3.

27 ACS Paragon Plus Environment

Environmental Science & Technology

808 809 810

811

Page 28 of 30

Figure 1. Top: Model predicted (line) and measured (diamonds) isoprene remaining in the chamber at steady state. Bottom: Measured C5H12O6 (blue x’s) and SOA (black squares) mass concentration. Note the different y-axis scale.

Figure 2. Measured mass yield of C5H12O6 (black x’s) and several model results with varying isomerization rates of C5H11O6 peroxy radical precursor as a function of input H2O2 (ppm). Inset provides a zoomed-in view to better observe the behavior of the measurements and most representative model run. Measurement error bars reflect our uncertainty in the calibration factor. 28 ACS Paragon Plus Environment

Page 29 of 30

812 813 814 815 816 817

818

Environmental Science & Technology

Scheme 3. Proposed detailed mechanism and calculated rates of the isomerization of one isomer of C5H11O6, the peroxy radical formed from ISOPOOH + OH in scheme 1.

Figure 3. The ratio of the sum of C5H11O6• isomerization products relative to the product of reaction with HO2 for measurements (black x’s) and various model results (lines) at varying C5H11O6• isomerization rates. Measurement error bars reflect our uncertainty in calibration factors.

29 ACS Paragon Plus Environment

Environmental Science & Technology

Page 30 of 30

Figure 4. Top: model OA concentrations plotted versus -5 -1 measurements for model runs with 10 s wall loss, 11% -1

C5H11O6 yield, and varying isomerization rates (s ). Bottom: pie charts of the most prolific particle phase species modeled (left) for the data point in the above plot -3

819 820 821 822 823 824 825 826 827 828 829 830 831 832 833

circled in yellow, and measured (right) in µg m .

TOC Graphic

834

30 ACS Paragon Plus Environment