Identification and Quantification of 4-Nitrocatechol ... - ACS Publications


Identification and Quantification of 4-Nitrocatechol...

0 downloads 83 Views 762KB Size

Subscriber access provided by READING UNIV

Article

Identification and Quantification of 4-Nitrocatechol Formed from OH and NO Radical-Initiated Reactions of Catechol in Air in the Presence of NO: Implications for Secondary Organic Aerosol Formation from Biomass Burning 3

x

Zachary Finewax, Joost A. de Gouw, and Paul J. Ziemann Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b05864 • Publication Date (Web): 22 Jan 2018 Downloaded from http://pubs.acs.org on January 24, 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 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 33

Environmental Science & Technology

1 2 3 4 5

Identification and Quantification of 4-Nitrocatechol Formed from OH and NO3

6

Radical-Initiated Reactions of Catechol in Air in the Presence of NOx:

7

Implications for Secondary Organic Aerosol Formation from Biomass Burning

8

Zachary Finewax†, §, Joost A. de Gouw§, ζ, Paul J. Ziemann†, § *

9 10 11 12 13 14 15 16 17 18 19

Submitted to Environmental Science and Technology

20 21 22



23

United States

24

§

25

Boulder, Colorado 80309, United States

26

ζ

27

80305, United States

28

*

29

Telephone: 303-492-9654

30

Fax: 303-492-1149

31

Email: [email protected]

Department of Chemistry and Biochemistry, University of Colorado, Boulder, Colorado 80309,

Cooperative Institute for Research in Environmental Sciences (CIRES), University of Colorado,

Chemical Sciences Division, NOAA Earth System Research Laboratory, Boulder, Colorado

Author to whom correspondence should be addressed.

1

ACS Paragon Plus Environment

Environmental Science & Technology

32 33

ABSTRACT Catechol (1,2-benzenediol) is emitted from biomass burning and produced from reaction

34

of phenol with OH radicals. It has been suggested as an important secondary organic aerosol

35

(SOA) precursor, but the mechanisms of gas-phase oxidation and SOA formation have not been

36

investigated in detail. In this study, catechol was reacted with OH and NO3 radicals in the

37

presence of NOx in an environmental chamber to simulate daytime and nighttime chemistry.

38

These reactions produced SOA with exceptionally high mass yields of 1.34 ± 0.20 and 1.50 ±

39

0.20, respectively, reflecting the low volatility and high density of reaction products. The

40

dominant SOA product, 4-nitrocatechol, for which an authentic standard is available, was

41

identified through thermal desorption particle beam mass spectrometry and Fourier transform

42

infrared spectroscopy, and was quantified in filter samples by liquid chromatography using UV

43

detection. Molar yields of 4-nitrocatechol were 0.30 ± 0.03 and 0.91 ± 0.06 for reactions with

44

OH and NO3 radicals, and thermal desorption measurements of volatility indicate that it is semi-

45

volatile at typical atmospheric aerosol loadings, consistent with field studies that have observed

46

it in aerosol particles. Formation of 4-nitrocatechol is initiated by abstraction of a phenolic H-

47

atom by an OH or NO3 radical to form a β-hydroxyphenoxy/o-semiquinone radical, which then

48

reacts with NO2 to form the final product.

49 50 51

INTRODUCTION Biomass burning is a significant source of atmospheric gas- and particle-phase emissions

52

worldwide.1 Biomass burning can be anthropogenic (heating, cooking, prescribed forest fire

53

burns) or natural (lightning propagated wildfires).2 The aerosol emitted from biomass burning

54

can impact regional air quality and regional and global climate through long-range transport,

2

ACS Paragon Plus Environment

Page 2 of 33

Page 3 of 33

Environmental Science & Technology

55

depending on the strength of convection in the plume.3 In addition to the primary organic aerosol

56

(POA) that is directly emitted from wildfires, secondary organic aerosol (SOA) can be formed

57

when emissions of volatile organic compounds (VOCs) are oxidized to condensable products.

58

Field studies have shown mixed results for O3 and aerosol formation in wildfire plumes, with

59

some studies indicating formation and others showing net losses.4-6 Furthermore, POA emissions

60

and SOA production from biomass burning are highly variable and depend on the type of fuel

61

burned.4 These findings show that the mechanisms of SOA formation from wildfire emissions

62

are complex, and may be difficult to model.

63

To date, a number of studies have been conducted to identify and quantify the complex

64

mixture of VOCs emitted from combustion of different types of biomass,7-9 but the fate of these

65

compounds with respect to reactions in the atmosphere and SOA formation is much less certain.

66

In one approach, Gilman et al.9 used their measurements of VOCs emitted from controlled

67

biomass burns with compound-specific SOA formation potentials (derived by Derwent et al.10

68

using the Master Chemical Mechanism11) to estimate total SOA formation potentials for

69

individual compounds and various compound classes. In light of unexplained enhancements and

70

suppressions of O3, and organic aerosol observed in numerous wildfire plumes, however, it is

71

apparent that the chemistry occurring within these plumes needs to be systematically investigated

72

to better constrain the processes leading to these observations.

73

1,2-Benzenediol, commonly known as catechol, is emitted directly from biomass

74

burning12 as well as produced in the atmosphere through the gas-phase reaction of benzene and

75

phenol with OH radicals.13,14 The emission of catechol from wildfires is due to pyrolysis of

76

lignin, a polymer in wood containing phenolic moieties.15 Catechol displays appreciable gas-

77

phase reactivity. Using average concentrations (molecules cm-3) of 2 × 106 (0.1 ppt), 7 × 1011 (30

3

ACS Paragon Plus Environment

Environmental Science & Technology

78

ppb), and 5 × 108 (20 ppt) for OH (12-h daytime), O3 (24-h), and NO3 (12-h nighttime),

79

respectively,16 and corresponding reaction rate constants of 1.0 × 10-10, 9.6 × 10-18, and 9.8 × 10-

80

11

81

studies have shown that catechol produces large quantities of SOA when reacted with O3 and OH

82

radicals,20,21 and it has been predicted to be one of the main precursors to SOA formation from

83

wildfire emissions.9 A study of gas-phase reaction products has shown that under moderate NOx

84

conditions, reaction with OH radicals forms significant amounts of 4-nitrocatechol (4NC),21 and

85

the Master Chemical Mechanism (MCM v3.1) assumes that 4NC is the sole reaction product.11

86

To our knowledge, however, the yield of 4NC has not been measured for reactions of catechol

87

with OH or NO3 radicals in the presence of NOx.

88

cm3 molecule-1 s-1,17-19 estimated atmospheric lifetimes are 1.4 h, 1.7 d, and 20 s. Chamber

In this study, catechol was reacted with OH and NO3 radicals in the presence of NOx to

89

determine yields of SOA and reaction products as well as reaction mechanisms under simulated

90

daytime and nighttime conditions. The vapor pressure of 4NC, the major SOA product in these

91

reactions, was measured for use in predicting its fate and possible environmental impacts.

92

Because of the large catechol emission factors for combustion of biomass,12 and its relatively

93

high predicted SOA formation potential,9 catechol may be responsible for an appreciable portion

94

of the SOA produced from wildfires.

95

MATERIALS AND METHODS

96

Chemicals. The chemicals used, with purities or solvent grades and supplier were

97

methanol, ethyl acetate, and acetonitrile (HPLC grade, EMD Millipore); water (HPLC grade,

98

Fisher Chemical); catechol (+99%, Sigma Aldrich); 4-nitrocatechol (97%, Sigma Aldrich);

99

glacial acetic acid (99.7%, Macron); dioctyl sebacate (>97%, Fluka); and nitric oxide (99%,

100

Matheson Tri-gas). Methyl nitrite and N2O5 were synthesized according to the procedures of

4

ACS Paragon Plus Environment

Page 4 of 33

Page 5 of 33

Environmental Science & Technology

101

Taylor et al.22 and Atkinson et al,23 respectively. Methyl nitrite was stored in a lecture bottle on a

102

vacuum manifold until used, and N2O5 was stored in a –80°C freezer and placed on a vacuum

103

manifold when used.

104

Environmental Chamber Experiments. Catechol reactions were conducted at 25 °C

105

and 630 Torr in an 8.0 m3 FEP Teflon environmental chamber filled with dry, clean air (RH <

106

1%, VOCs and NOx 96% of 4NC was in particles. In addition, we note

322

that although there are significant uncertainties in the quantities that contribute to this

323

calculation, if the true value was significantly less than 96%, then one would expect that the

324

measured 4NC yields would increase with increased aerosol mass concentration because of

325

enhanced partitioning of 4NC from the gas phase to the particles. But since the 4NC yields for

326

the OH radical-initiated reactions are essentially independent of SOA mass concentration from

327

300 to 2500 ug m-3 (the four experiments shown in Table 1), the 4NC appears to be almost

328

entirely in the particle phase in all experiments. This result thus supports our assumption that the

329

molar yield of 4NC determined from particle analysis is essentially equal to the total molar yield.

14

ACS Paragon Plus Environment

Page 15 of 33

Environmental Science & Technology

330

The molar yields of 4NC measured for reactions of catechol with OH and NO3 radicals were

331

0.30 ± 0.03 and 0.91 ± 0.06, respectively (Table 1). The much higher 4NC yield for the NO3

332

radical reaction is consistent with the much greater similarity in SOA mass spectra (Figure 1),

333

TPTD profiles (Figure S6), and ATR-FTIR spectra (Figure 2). Because of the nonlinear

334

dependence of ATR-FTIR spectral intensity on species concentration, due to film thickness

335

effects, no attempt was made to quantify the yield of 4NC in SOA by this method or to compare

336

the ATR-FTIR spectra quantitatively between 4NC and the filter extract. It is also worth noting

337

that although both reactions produced colored aerosol (Figure S7), SOA formed from the OH

338

radical reaction was orange/brown whereas SOA formed from the NO3 radical reaction was light

339

yellow: the color of 4NC.

340

One of the additional components of SOA formed from the OH radical reaction appears

341

to be 5-nitro-1,2,3-benzenetriol. The presence of this compound is indicated by the molecular ion

342

peak at m/z 171 in the TDPBMS mass spectrum (Figure 1A), which is absent from the mass

343

spectrum of SOA formed from the NO3 radical reaction (Figure 1B). Furthermore, when one of

344

the SOA components fractionated by HPLC (Figure S8)was analyzed by CI-ITMS, the mass

345

spectrum (Figure S9) contained peaks at m/z 172 [M + H]+, 155 [M + H – OH]+, 141 [M + H –

346

NO]+, 126 [M + H – NO2]+, and 108 [M + H – NO2 – H2O]+, which are all consistent with this

347

product. For nitroaromatic compounds, loss of NO and NO2 from the protonated molecular ion is

348

enhanced if electron donating groups are located meta to the –NO2 group, whereas loss of H2O is

349

enhanced when an electron withdrawing –OH group is located ortho to the –NO2 group.45

350

Because of the very low signal at m/z 154, which corresponds to the [M + H – H2O]+ ion that

351

would be formed by the latter pathway, the only arrangement of one –NO2 and three –OH groups

15

ACS Paragon Plus Environment

Environmental Science & Technology

352

on the ring that remains possible is 5-nitro-1,2,3-benzenetriol. The measured molar yield of 5-

353

nitro-1,2,3-benzenetriol was 0.04 ± 0.01.

354

The sum of the mass yields (mass of product formed/mass of catechol reacted) of 4NC

355

(0.42) and 5-nitro-1,2,3-benzenetriol (0.06) formed in the OH radical reaction was 0.48, which is

356

much lower than the mass yield of SOA (see discussion below) of 1.34, and indicates that only

357

~35% of the SOA products are accounted for in the molecular analysis. In an attempt to further

358

characterize the unidentified components of SOA formed in the OH radical reaction, the SOA

359

was reacted with 2,4-dinitrophenylhydrazine (DNPH) to convert compounds containing carbonyl

360

groups to hydrazones that can be detected by their UV-Vis absorbance.33,46 The ATR-FTIR

361

spectrum of the underivatized SOA (Figure 2) indicated that compounds with carbonyl groups

362

were present, due to absorbance at 1600–1700 cm-1, and would be indicative of ring-opened

363

products. The C-H stretching at 2900 – 3000 cm-1 may be additional evidence of ring-opened

364

products, but is largely from DOS seed particles. Peaks in HPLC chromatograms (Figure S9)

365

obtained at 360 nm and 430 nm indicated the presence of monofunctional and difunctional

366

carbonyls,46 but CI-ITMS analysis was unsuccessful at identifying these compounds due to the

367

complex mass spectra. Based on these results, it is likely that the SOA contained hemiacetals and

368

acetals formed from reactions involving ring-opened products. This is supported by the

369

observation of broad peaks from ~60–150oC in the TPTD profiles of total ion signal

370

(proportional to aerosol mass) beyond the 4NC peak at 39oC (Figure S6A).

371

SOA Yields. The yields of SOA formed from reactions of catechol with OH and NO3

372

radicals were 1.34 ± 0.2 and 1.50 ± 0.2, with uncertainties reported as the standard deviation of 4

373

and 2 replicate experiments, respectively, and are given in Table 1. These values are

374

exceptionally high, considering that those measured for the photooxidation of simple aromatics

16

ACS Paragon Plus Environment

Page 16 of 33

Page 17 of 33

Environmental Science & Technology

375

ranged from about 0 to 0.14,47 whereas those measured for reactions of biogenic compounds with

376

OH and NO3 radicals and O3 ranged from 0 to 0.9.48 Yields can be larger than 1.0 because they

377

are based on the mass of SOA formed relative to the mass of catechol reacted, and the SOA

378

products contain mass from O and N that was not associated with catechol. Note also that the

379

yield of 1.50 for the reaction with NO3 radicals is equivalent to a molar yield of 4NC of 1.06

380

(assuming the SOA is pure 4NC), consistent within measurement uncertainties with the value of

381

0.91 measured by HPLC. Although the SOA yield for the reaction with NO3 radicals is 12%

382

larger than for the reaction with OH radicals, the 4NC yield is larger by a factor of 3. This shows

383

again that contributions from products other than 4NC comprise a considerable fraction of the

384

SOA formed from the OH radical reaction, whereas reaction with NO3 quantitatively produces

385

4NC.

386

The SOA yield of 1.34 measured here for the OH radical-initiated reaction is much

387

higher than the value of 0.53 reported previously by Borrás and Tortajada-Genaro.21 This is most

388

likely due to the loss of fewer reaction products to the chamber walls during our 5 min (the time

389

that the lights were turned on to create OH radicals) experiments compared to their 7 h

390

experiments, and also our use of seed particles, whose surface area enhanced the rate of

391

condensation of reaction products to particles and thus reduced loses due to condensation on the

392

chamber walls.24,25 The experiments conducted by Borrás and Tortajada-Genaro21 also differed

393

from ours in that they allowed O3 to form (by not adding NO), and O3 could compete with OH

394

radicals in reactions with catechol.17,18

395

ATMOSPHERIC IMPLICATIONS

396 397

The yields of 4NC measured here for the reaction of catechol with OH and NO3 radicals in the presence of NOx have implications for explicit chemical modeling. In MCM v3.1, the

17

ACS Paragon Plus Environment

Environmental Science & Technology

398

reaction of catechol with OH radicals is assumed to produce 4NC with unit yield.11 Our results

399

suggest, however, that although this is valid for the reaction with NO3 radicals, other products

400

are formed in significant yields in the OH radical reaction. Further identification and

401

quantification of those products are needed to determine other reaction pathways, with an

402

important question being the extent to which abstraction of a phenolic hydrogen competes with

403

addition of an OH radical to the aromatic ring.

404

By considering various potential loss processes and the properties of 4NC, one can obtain

405

a sense of its fate following formation from reaction of catechol with OH or NO3 radicals near a

406

wildfire plume. The short atmospheric lifetimes of catechol (estimated above as approximately

407

1.4 h and 20 s for reactions with OH and NO3 radicals at global average oxidant concentrations)

408

and the high concentrations of NO2 present in the early stages of a wildfire plume suggests rapid

409

removal of catechol and rapid 4NC formation close to the wildfire. Although the rate constants

410

for reaction of 4NC with OH and NO3 radicals have not been measured, it is known that,

411

compared to their unsubstituted aromatic analogs, nitroaromatics react ~30 times more slowly

412

with OH radicals49,50 and >100 times more slowly with NO3 radicals.39 Assuming a similar

413

relationship for catechol and 4NC, the lifetimes of 4NC with respect to gas-phase reactions with

414

OH and NO3 radicals are expected to be about 42 h and 30 min. The overall lifetimes are likely

415

to be longer than this, however, since the saturation concentration of 4NC (~13 µg m-3) indicates

416

that it should be present in both the gas and organic particle phases in typical ambient

417

environments, where organic aerosol loadings are ~1–10 µg m-3,51 with a higher fraction in

418

particles near wildfires, where loadings can reach ~100 µg m-3.52 This is consistent with the

419

detection of 4NC and its methylated derivatives in aerosol sampled from biomass burning events

420

during winter at concentrations of 0.1–100 ng m-3, 53–55 with cold temperatures reducing the

18

ACS Paragon Plus Environment

Page 18 of 33

Page 19 of 33

Environmental Science & Technology

421

vapor pressure and enhancing gas-to-particle partitioning of 4NC. Furthermore, although the

422

Henry’s Law constant for 4NC has not been reported, by comparing values for phenol, 4-

423

nitrophenol, and catechol (2.8 x 103,56 2.1 x 104,56 and 8.3 x 105 mol L-1 atm-1,57 respectively), it

424

is expected that the value for 4NC will be at least 106 mol L-1 atm-1. The high water solubility of

425

4NC thus allows it to be taken up into aqueous particles,58 which is consistent with its observed

426

presence in rainwater.58 Since a considerable fraction of 4NC should be present in the organic

427

and aqueous phases of aerosol particles, which will reduce the amount available for removal by

428

reaction with OH and NO3 radicals, it is possible that 4NC can undergo long-range transport. It is

429

also worth noting that although this study was concerned with gas-phase reactions of catechol,

430

uptake into aqueous particles containing dissolved H2O2 and iron can lead to Fenton chemistry,59,

431

60

432

can occur in the presence and absence of light suggests that there are additional competing

433

processes to the gas-phase formation of 4NC from reaction of catechol with OH and NO3

434

radicals.

which mainly proceeds by hydroxylation of the ring to produce benzenetriols.59, 60 That this

435

The results from this work can also be used to provide insight into the possible

436

importance of catechol oxidation to SOA formed from biomass burning emissions. Using the

437

SOA yield of 1.34 measured for the reaction of catechol with OH radicals, and an emission ratio

438

of 4 mmol catechol mol-1 CO,9,12 we estimate an organic aerosol enhancement of 23 μg m-3

439

ppmv-1 CO over background concentrations. It should be emphasized, however, that the emission

440

ratio of catechol used in this estimate exhibits large variability between fires, burn conditions,

441

and biomass fuel type.9,12 Also, because the dominant OH radical reaction product, 4NC, is

442

expected to exist less fully in the particle phase in the atmosphere than in the chamber

443

experiments, the SOA yield of 1.34 may be too high. Nonetheless, when compared to POA and

19

ACS Paragon Plus Environment

Environmental Science & Technology

444

SOA concentrations reported for biomass burning field campaigns,1 this estimate suggests that

445

catechol oxidation alone could account for ~10–40% of the organic aerosol mass (see Figure 1B

446

in de Gouw and Jimenez1). This agrees reasonably well with the prediction of Veres et al.,12

447

where catechol emission ratios were measured by burning wood in the laboratory. Furthermore,

448

in a series of wood burning experiments, Bruns et al.61 measured the amounts of VOCs reacted

449

and SOA formed when emissions were reacted with OH radicals generated by photolysis of

450

HONO. By multiplying the amount of VOCs reacted by published SOA yields they estimated

451

that 22 VOCs could account for 84–116% of measured SOA mass, with 8% being due to

452

catechol oxidation. If our measured SOA yield of 1.34 is used instead of the value of 0.39 taken

453

from the literature,62 then catechol oxidation would account for 27% of SOA mass. This may be

454

a more accurate estimate than that of Bruns et al.,61 since the SOA yield of 0.39 was measured

455

for catechol oxidation in the absence of NOx,62 conditions that probably do not represent the

456

chemistry that occurred during their HONO photolysis experiments. Finally, the mechanism of formation and optical properties of 4NC, as well as

457 458

methylnitrocatechols, suggests a possible role for these compounds in O3 formation in biomass

459

burning plumes. Iinuma et al.53 have shown that methylnitrocatechols are produced from the

460

reaction of methylhydroxybenzenes with OH radicals (analogous to formation of 4NC from

461

catechol), but with an estimated molar yield in the aerosol of 0.003, and have suggested them as

462

aerosol tracers of biomass burning. Nitroaromatic compounds are components of brown carbon63,

463

64

464

thus attenuate UV light.66 This attenuation, and the removal of NOx by formation of

465

nitroaromatics, may play a role in the lack of observed O3 formation in some wildfire plumes.6

466

For example, NOx emission ratios average 30 ppbv ppmv-1 CO for pine fuels,67 while catechol

(as shown from the filters in Figure S7 and absorbance measurements by Hinrichs et al.65), and

20

ACS Paragon Plus Environment

Page 20 of 33

Page 21 of 33

Environmental Science & Technology

467

emission ratios can vary from 0.2 – 4 ppbv ppmv-1 CO.9,12 When combined with our results of a

468

30% yield of 4NC from catechol oxidation by OH in the presence of NOx, this suggests that up

469

to 1.2 ppbv [ppmv-1 CO]-1 of NOx could be lost as a result of 4NC formation, representing 4% of

470

emitted NOx. The sequestration of NOx can also occur through formation of peroxyacyl nitrates

471

that can be efficiently formed from alkenes and acetaldehyde, which are all abundant in wildfire

472

smoke.

473

ACKNOWLEDGMENTS

474

This work was funded by the Department of Energy, Office of Biological and

475

Environmental Sciences under Grant DE-SC0010470, and by the National Science Foundation

476

under Grant AGS-1420007. Any opinions, findings, and conclusions or recommendations

477

expressed in this material are those of the author and do not necessarily reflect the views of the

478

National Science Foundation (NSF).

21

ACS Paragon Plus Environment

Environmental Science & Technology

479 480

Page 22 of 33

Table 1: Experimental conditions and yields of SOA and 4NC formed from reactions of catechol (Cat) with OH and NO3 radicals.

481 482

a

483 484 485 486

b

Experiment

Cata (ppb)

∆Cata (ppb)

CH3ONO (ppb)

NO (ppb)

JNO2 (min-1)

N 2O 5 (ppb)

SOA mass yield

4NCb molar yield

Catechol + OH

95 ± 5

57 ± 5

1030

1010

0.14

0

1.40

0.31 ± 0.03

Catechol + OH

670 ± 40

560 ± 40

4940

4980

0.37

0

1.11

0.33 ± 0.03

Catechol + OH

140 ± 20

100 ± 30

4060

4110

0.14

0

1.45

0.28 ± 0.08

Catechol + OH

50 ± 5

29 ± 5

1020

5090

0.37

0

1.38

0.27 ± 0.06

Catechol + NO3

193 ± 8

124 ± 9

0

0

0

320

1.61

0.95 ± 0.08

Catechol + NO3

185 ± 10

150 ± 14

0

0

0

480

1.38

0.86 ± 0.05

Uncertainties for catechol concentrations are standard deviations calculated from the duplicate measurements made before and after reaction. Uncertainties in 4NC yield were calculated by propagation of uncertainties in catechol concentration, flow rates for filter sampling, mass measurements, and the 4NC HPLC calibration curve.

22

ACS Paragon Plus Environment

Page 23 of 33

Environmental Science & Technology

487 488 489

Figure 1. Real-time TDPBMS mass spectra of SOA formed from reactions of catechol with (A) OH radicals, and (B) NO3 radicals, and (C) aerosol formed from atomization of a 4NC standard. 23

ACS Paragon Plus Environment

Environmental Science & Technology

490

491 492 493 494 495 496 497

Figure 2. ATR-FTIR spectra of SOA formed from reactions of catechol with (A) OH radicals and (B) NO3 radicals, and (C) a 4NC standard. Wavenumber ranges correspond to the following functional group vibrational modes: O-H stretch (3500 – 3100 cm-1), C-H stretch from DOS (3000-2900 cm-1), carbonyl C=O stretch (1750-1650 cm-1), aromatic C=C stretch (1600 – 1450 cm-1), and nitro N-O stretch (1330 – 1280 cm-1).

24

ACS Paragon Plus Environment

Page 24 of 33

Page 25 of 33

Environmental Science & Technology

498

499 500 501

Figure 3. Proposed mechanism for forming 4NC from reactions of catechol with OH or NO3 radicals in air in the presence of NOx.

25

ACS Paragon Plus Environment

Environmental Science & Technology

502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546

SUPPORTING INFORMATION The supporting information contains the following information: HPLC calibration curves of 4NC and 5-nitro-1,2,3-benzenetriol; UV-Vis spectra of 4NC and 5-nitro-1,2,3-benzenetriol; proposed electron ionization mechanism for 4NC; thermal desorption profiles of SOA and 4NC; photograph of SOA collected on Teflon filters; HPLC chromatogram of SOA; CI-ITMS mass spectra of 5-nitro-1,2,3-benzenetriol; HPLC chromatogram of carbonyl derivatized SOA; and references. REFERENCES 1. de Gouw, J. A.; Jimenez, J. L. Organic aerosols in the Earth’s atmosphere. Environ. Sci. Technol. 2009, 43, 7614–7618. 2. Crutzen, P. J.; Andreae, M. O. Biomass burning in the tropics: impact on atmospheric chemistry and biogeochemical cycles. Science 1990, 250, 1669–1678. 3. Val Martin, M.; Logan, J. A.; Kahn, R. A.; Leung, F. -Y.; Nelson, D. L.; Diner, D. J. Smoke injection heights from fires in North America: analysis of 5 years of satellite observations. Atmos. Chem. Phys. 2010, 10, 1491–1510. 4. Ortega, A. M.; Day, D. A.; Cubison, M. J.; Brune, W. H.; Bon, D.; de Gouw, J. A.; Jimenez, J. L. Secondary organic aerosol formation and primary organic aerosol oxidation from biomassburning smoke in a flow reactor during FLAME-3. Atmos. Chem. Phys. 2013, 13, 11551–11571. 5. Hecobian, A.; Liu, Z.; Hennigan, C. J.; Huey, L. G.; Jimenez, J. L.; Cubison, M. J.; Vay, S.; Diskin, G. S.; Sachse, G. W.; Wisthaler, A.; Mikoviny, T.; Weinheimer, A. J.; Liao, J.; Knapp, D. J.; Wennberg, P. O.; Kürten, A.; Crounse, J. D.; St. Clair, J.; Wang, Y.; Weber, R. J. Comparison of chemical characteristics of 495 biomass burning plumes intercepted by the NASA DC-8 aircraft during the ARCTAS/CARB-2008 field campaign. Atmos. Chem. Phys. 2011, 11, 13325—13337. 6. Jaffe, D. A.; Wigder, N. L. Ozone production from wildfires: a critical review. Atmos. Environ. 2012, 51, 1–10. 7. Yokelson, R. J.; Burling, I. R.; Gilman, J. B.; Warneke, C.; Stockwell, C. E.; de Gouw, J.; Akagi, S. K.; Urbanski, S. P.; Veres, P.; Roberts, J. M.; Kuster, W. C.; Reardon, J.; Griffith, D. W. T.; Johnson, T. J.; Hosseini, S.; Miller, J. W.; Cocker III, D. R.; Jung, H.; Weise, D. R. Coupling field and laboratory measurements to estimate the emission factors of identified and unidentified trace gases for prescribed fires. Atmos. Chem. Phys. 2013, 13, 89–116. 8. Stockwell, C. E.; Veres, P. R.; Williams, J.; Yokelson, R. J. Characterization of biomass burning emissions from cooking fires, peat, crop residue, and other fuels with high-resolution proton-transfer-reaction time-of-flight mass spectrometry. Atmos. Chem. Phys. 2015, 15, 845– 865.

26

ACS Paragon Plus Environment

Page 26 of 33

Page 27 of 33

547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591

Environmental Science & Technology

9. Gilman, J. B.; Lerner, B. M.; Kuster, W. C.; Goldan, P. D.; Warneke, C.; Veres, P. R.; Roberts, J. M.; de Gouw, J. A.; Burling, I. R.; Yokelson, R. J. Biomass burning emissions and potential air quality impacts of volatile organic compounds and other trace gases from fuels common in the US. Atmos. Chem. Phys. 2015, 15, 13915–13938. 10. Derwent, R. G.; Jenkin, M. E.; Utembe, S. R.; Shallcross, D. E.; Murrells, T. P.; Passant, N. R. Secondary organic aerosol formation from a large number of reactive manmade organic compounds. Sci. Total Environ. 2010, 408, 3374–3381. 11. Bloss, C.; Wagner, V.; Jenkin, M. E.; Volkamer, R.; Bloss, W. J.; Lee, J. D.; Heard, D. E.; Wirtz, K.; Martin-Reviejo, M.; Rea, G.; Wenger, J. C.; Pilling, M. J. Development of a detailed chemical mechanism (MCMv3.1) for the atmospheric oxidation of aromatic hydrocarbons. Atmos. Chem. Phys. 2005, 5, 641–664. 12. Veres, P.; Roberts, J. M.; Burling, I. R.; Warneke, C.; de Gouw, J.; Yokelson, R. J. Measurements of gas‐phase inorganic and organic acids from biomass fires by negative‐ion proton‐transfer chemical‐ionization mass spectrometry. J. Geophys. Res. 2010, 115, D23302. 13. Olariu, R. I.; Klotz, B.; Barnes, I.; Becker, K. H.; Mocanu, R. FT-IR study of the ringretaining products from the reaction of OH radicals with phenol, o-, m-, and p-cresol. Atmos. Environ. 2002, 36, 3685–3697. 14. Volkamer, R.; Klotz, B.; Barnes, I.; Imamura, T.; Wirtz, K.; Washida, N.; Becker, K. H.; Platt, U. OH-initiated oxidation of benzene part I. phenol formation under atmospheric conditions. Phys. Chem. Chem. Phys. 2002, 4, 1598–1610. 15. Simoneit, B. R. T.; Rogge, W. F.; Mazurek, M. A.; Standley, L. J.; Hildemann, L. M.; Cass, G. R. Lignin pyrolysis products, lignans, and resin acids as specific tracers of plant classes in emissions from biomass combustion. Environ. Sci. Technol. 1993, 27, 2533–2541. 16. Atkinson, R.; Arey, J. Atmospheric degradation of volatile organic compounds. Chem. Rev. 2003, 103, 4605−4638. 17. Olariu, R. I.; Barnes, I.; Becker, K. H., Klotz, B. Rate coefficients for the gas-phase reaction of OH radicals with selected dihydroxybenzenes and benzoquinones. Int. J. Chem. Kinet. 2000, 32, 696–702. 18. Tomas, A.; Olariu, R. I.; Barnes, I.; Becker, K. H. Kinetics of the reactions of O3 with selected benzenediols. Int. J. Chem. Kinet. 2003, 6, 223–230. 19. Olariu, R. I.; Tomas, A.; Barnes, I.; Bejan, I.; Becker, K. H.; Wirtz, K. Rate coefficients for the gas-phase reaction of NO3 radicals with selected dihydroxybenzenes. Int. J. Chem. Kinet. 2004, 36, 577–583. 20. Coeur-Tourneur, C.; Tomas, A.; Guilloteau, A.; Henry, F.; Ledoux, F.; Visez, N.; Riffault,

27

ACS Paragon Plus Environment

Environmental Science & Technology

592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635

V.; Wenger, J. C.; Bedjanian, Y. Aerosol formation yields from the reaction of catechol with ozone. Atmos. Environ. 2009, 43, 2360–2365. 21. Borrás, E.; Tortajada-Genaro, L. A. Secondary organic aerosol formation from the photooxidation of benzene. Atmos. Environ. 2012, 47, 154–163. 22. Taylor, W. D.; Allston, T. D.; Moscato, M. J.; Fazekas, G. B.; Kozlowski, R.; Takacs, G. A. Atmospheric photodissociation lifetimes for nitromethane, methyl nitrite, and methyl nitrate. Int. J. Chem. Kinet. 1980, 12, 231–240. 23. Atkinson, R.; Plum, C. N.; Carter, W. P. L.; Winer, A. M.; Pitts, J. N., Jr. Rate Constants for the Gas-Phase Reactions of Nitrate Radicals with a Series of Organics in Air at 298 ± 1 K. J. Phys. Chem. 1984, 88, 1210−1215. 24. 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, 881–892. 25. 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, 5757–5765. 26. Atkinson, R.; Carter, W. P.; Winer, A. M.; Pitts, J. N., Jr. An experimental protocol for the determination of OH radical rate constants with organics using methyl nitrite photolysis as an OH radical source. J. Air Pollut. Control Assoc. 1981, 31, 1090−1092. 27. Finlayson-Pitts, B. J., Pitts, J. The Chemistry of the Upper and Lower Atmosphere; Academic Press; USA, 2000. 28. Boublik, T.; Fried, V.; Hala, E. The vapour pressures of pure substances, 2nd, ed.; Elsevier: Amsterdam, 1984. 29. Tobias, H. J.; Kooiman, P. M.; Docherty, K. S.; Ziemann, P. J. Real-time chemical analysis of organic aerosols using a thermal desorption particle beam mass spectrometer. Aerosol Sci. Technol. 2000, 33, 170–190. 30. Tobias, H. J.; Ziemann, P. J. Compound identification in organic aerosols using temperatureprogrammed thermal desorption particle beam mass spectrometry. Anal. Chem. 1999, 71, 3428– 3435. 31. Docherty, K. S.; Ziemann, P. J. Effects of stabilized Criegee intermediate and OH radical scavengers on aerosol formation from reactions of α-pinene with O3. Aerosol Sci. Technol. 2003, 37, 877–891.

28

ACS Paragon Plus Environment

Page 28 of 33

Page 29 of 33

636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681

Environmental Science & Technology

32. Matsunaga, A.; Ziemann, P. J. Yields of β-hydroxynitrates and dihydroxynitrates in aerosol formed from OH radical-initiated reactions of linear alkenes in the presence of NOx. J. Phys. Chem. A 2009, 113, 599–606. 33. Ranney, A. P.; Ziemann, P. J. Identification and quantification of oxidized organic aerosol compounds using derivatization, liquid chromatography, and chemical ionization mass spectrometry. Aerosol Sci. Technol. 2017, 51, 342–353. 34. Yeh, G. K.; Claflin, M. S.; Ziemann, P. J. Products and mechanism of the reaction of 1pentadecene with NO3 radicals and the effect of a –ONO2 group on alkoxy radical decomposition. J. Phys. Chem. A 2015, 43, 10684–10696. 35. Yeh, G. K.; Ziemann, P. J. Alkyl nitrate formation from the reactions of C8-C14 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, 8147–8157. 36. http://webbook.nist.gov/cgi/cbook.cgi?ID=C3316094&Units=SI&Mask=200#Mass-Spec 37. Yinon, J. Mass spectral fragmentation pathways in some dinitroaromatic compounds studied by collision-induced dissociation and tandem mass spectrometry. Org. Mass. Spectrom. 1992, 27, 689–694. 38. McLafferty, F. W., Tureček, F. Interpretation of Mass Spectra, 4th, ed.; University Science Books: USA, 1993. 39. Atkinson, R.; Aschmann, S. M.; Arey, J. Reactions of OH and NO3 radicals with phenol, cresols and 2-nitrophenol at 296 ± 2 K. Environ. Sci. Technol. 1992, 26, 1397–1403. 40. Nishino, N.; Arey, J.; Atkinson, R. Formation of glyoxal and methylglyoxal from the OH radical-initiated reaction of toluene, xylenes, and trimethylbenzenes as a function of NO2 concentration. J. Phys. Chem. A 2010, 114, 10140–10147. 41. Platz, J.; Nielsen, O. J.; Wallington, T. J.; Ball, J. C.; Hurley, M. D.; Straccia, A. M.; Schneider, W. F.; Sehested, J. Atmospheric chemistry of the phenoxy radical, C6H5O(•): UV spectrum and kinetics of its reaction with NO, NO2, and O2. J. Phys. Chem. A 1998, 102, 7964– 7974. 42. Niki, H.; Maker, P. D.; Savage, C. M.; Breitenbach, L. P. Fourier transform infrared (FTIR) studies of gaseous and particulate nitrogenous compounds, Nitrogenous Air Pollutants, Ann Arbor Science, Ann Arbor, MI, USA; 1979. 43. Pankow, J. F. An absorption model of gas/particle partitioning of organic compounds in the atmosphere. Atmos. Environ. 1994, 28, 185–188. 44. Lim, Y. B.; Ziemann, P. J. Chemistry of secondary organic aerosol formation from OH radical-initiated reactions of linear, branched, and cyclic alkanes in the presence of NOx. Aerosol

29

ACS Paragon Plus Environment

Environmental Science & Technology

682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727

Sci. Technol., 2009, 43, 604—619. 45. Harrison, A. G.; Kallury, R. K. M. R. The chemical ionization mass spectra of mononitroarenes. Org. Mass Spectrom. 1980, 15, 284–288. 46. Grosjean, E.; Green, P. G.; Grosjean, D. Liquid chromatography analysis of carbonyl (2,4dinitrophenyl)hydrazones with detection by diode array ultraviolet spectroscopy and by atmospheric pressure negative chemical ionization mass spectrometry. Anal. Chem. 1999, 71, 1851–1861. 47. Odum, J. R.; Jungkamp, T. P. W.; Griffin, R. J.; Forstner, H. J. L.; Flagan, R. C.; Seinfeld, J. H. Aromatics, reformulated gasoline, and atmospheric organic aerosol formation. Environ. Sci. Technol. 1997, 31, 1890–1897. 48. Griffin, R. J.; Cocker, D. R.; Flagan, R. C.; Seinfeld, J. H. Organic aerosol formation from the oxidation of biogenic hydrocarbons. J. Geophys. Res. 1999, 104, 3555–3567. 49. Zazo, J. A.; Casas, J. A.; Mohedano, A. F.; Gilarranz, M. A.; Rodriguez, J. J. Chemical pathway and kinetics of phenol oxidation by Fenton’s reagent. Environ. Sci. Technol. 2005, 39, 9295—9302. 50. Santos, P. S. M.; Duarte, A. C. Fenton-like oxidation of small aromatic acids from biomass burning in atmospheric water and in the absence of light: Identification for intermediates and reaction pathways. Chemosphere. 2016, 154, 599—603. 51. Atkinson, R.; Aschmann, S. M. Rate constants for the gas-phase reactions of the OH radical with the cresols and dimethylphenols at 296 +/- 2K. . Int. J. Chem. Kinet. 2000, 22, 59–67. 52. Bejan, I.; Barnes, I.; Olariu, R.; Zhou, S.; Wiesen, P.; Benter, T. Investigations on the gasphase photolysis and OH radical kinetics of methyl-2-nitrophenols. Phys. Chem. Chem. Phys. 2007, 9, 5686–5692. 53. Jimenez, J. L.; Jayne, J. T. Shi, Q.; Kolb, C. E.; Worsnop, D. R.; Yourshaw, I.; Seinfeld, J. H.; Flagan, R. C.; Zhang, X.; Smith, K. A.; Morris, J. W.; Davidovits, P. Ambient aerosol sampling using the Aerodyne Aerosol Mass Spectrometer. J. Geophys. Res., 2003, 108, doi:10.1029/2001JD001213. 54. Hobbs, P.V.; Reid, J.S.; Herring, J. A.; Nance, J. D.; Weiss, R. E.; Ross, J. L.; Hegg, D. A.; Ottmar, R. D.; Liousse, C. Particle and trace-gas measurements in the smoke from prescribed burns of forest products in the Pacific Northwest. Biomass Burning and Global Change. MIT Press, Cambridge, Mass., 1996. 55. Iinuma, Y.; Böge, O.; Gräfe, R.; Herrmann, H. Methyl-nitrocatechols: atmospheric tracer compounds for biomass burning secondary organic aerosols. Environ. Sci. Technol. 2010, 44, 8453–8459.

30

ACS Paragon Plus Environment

Page 30 of 33

Page 31 of 33

728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 770 771 772 773

Environmental Science & Technology

56. Gaston, C. J.; Lopez-Hilfiker, F. D.; Whybrew, L. E.; Hadley, O.; McNair, F.; Gao, H.; Jaffe, D. A.; Thornton, J. A. Online molecular characterization of fine particulate matter in Port Angeles, WA: evidence for a major impact from residential wood smoke. Atmos. Environ. 2016, 138, 99–107. 57. Mohr, C.; Lopez-Hilfiker, F. D.; Zotter, P.; Prévôt, A. S. H.; Xu, L.; Ng, N. L.; Herndon, S. C.; Williams, L. R.; Franklin, J. P.; Zahniser, M. S.; Worsnop, D. R.; Knighton, W. B.; Aiken, A. C.; Gorkowski, K. J.; Dubey, M. K.; Allan, J. D.; Thornton, J. A. Contribution of nitrated phenols to wood burning brown carbon light absorption in Detling, United Kingdom during winter time. Environ. Sci. Technol. 2013, 47, 6316–6324. 58. X. X. Guo; P. Brimblecombe. Henry’s law constants of phenol and mononitrophenols in water and aqueous sulfuric acid. Chemosphere, 2007, 68, 436—444. 59. HSDB: Hazardous Substances Data Bank, TOXicology data NETwork (TOXNET), National Library of Medicine (US), available at: http://toxnet.nlm.nih.gov/newtoxnet/hsdb.htm (last access: November 3, 2017), 2015. 60. Desyaterik, Y.; Sun, Y.; Shen, X.; Lee, T.; Wang, X.; Wang, T.; Collett, J. L. Speciation of “brown” carbon in cloud water impacted by agricultural biomass burning in eastern China. J. Geophys. Res. Atmos. 2013, 118, 1–11. 61. Bruns, E. A.; El Haddad, I.; Slowik, J. G.; Kilic, D.; Klein, F.; Baltensperger, U.; Prévôt, A. S. H. Identification of significant precursor gases of secondary organic aerosols from residential wood combustion. Sci. Rep., 2016, 6, DOI: 10.1038/srep27881. 62. Nakao, S.; Clark. C.; Tang, P.; Sato, K.; Cocker III, D. Secondary organic aerosol formation from phenolic compounds in the absence of NOx. Atmos. Chem. Phys., 2011, 11, 10649–10660. 63. Laskin, A.; Laskin, J.; Nizkorodov, S. A. Chemistry of atmospheric brown carbon. Chem. Rev. 2015, 115, 4335–4382. 64. Xie, M.; Chen, X.; Hayes, M. D.; Lewandowski, M.; Offenberg, J.; Kleindienst, T. E.; Holder, A. L. Light absorption of secondary organic aerosol: composition and contribution of nitroaromatic compounds. Environ. Sci. Technol. 2017, 51, 11607—11616. 65. Hinrichs, R. Z.; Buczek, P.; Trivedi, J. J. Solar absorption by aerosol-bound nitrophenols compared to aqueous and gaseous nitrophenols. Environ. Sci. Technol. 2016, 50, 5661–5667. 66. Chen, J.; Wenger, J. C.; Venables, D. S. Near-Ultraviolet absorption cross sections of nitrophenols and their potential influence on tropospheric oxidation chemistry. J. Phys. Chem. A 2011, 115, 12235–12242. 67. Yokelson, R. J.; Burling, I. R.; Gilman, J. B.; Warneke, C.; Stockwell, C. E.; de Gouw, J.; Akagi, S. K.; Urbanski, S. P.; Veres, P.; Roberts, J. M.; Kuster, W. C.; Reardon, J.; Griffith, D. W. T.; Johnson, T. J.; Hosseini, S.; Miller, J. W.; Cocker III, D. R.; Jung, H.; Weise, D. R.

31

ACS Paragon Plus Environment

Environmental Science & Technology

774 775 776 777 778 779 780 781

Coupling field and laboratory measurements to estimate the emission factors of identified and unidentified trace gases for prescribed fires. Atmos. Chem. Phys., 2013, 13, 89–116.

782 783 784 785 786 787 788 789 790 791 792 793 794 795 796 797 798 799 800 801 802 803 804 805 806 807

32

ACS Paragon Plus Environment

Page 32 of 33

Page 33 of 33

808

Environmental Science & Technology

TOC

809

33

ACS Paragon Plus Environment