Kinetics Study of OH Uptake onto Deliquesced NaCl Particles by

Particles by Combining Laser Photolysis and Laser- ... uptake coefficient for the atmospheric relevant OH concentration within the range from 0.77 to ...
2 downloads 0 Views 357KB Size
Subscriber access provided by Technical University of Munich University Library

Clusters, Radicals, and Ions; Environmental Chemistry

Kinetics Study of OH Uptake onto Deliquesced NaCl Particles by Combining Laser Photolysis and Laser-Induced Fluorescence Yosuke Sakamoto, Jun Zhou, Nanase Kohno, Maho Nakagawa, Jun Hirokawa, and Yoshizumi Kajii J. Phys. Chem. Lett., Just Accepted Manuscript • DOI: 10.1021/acs.jpclett.8b01725 • Publication Date (Web): 06 Jul 2018 Downloaded from http://pubs.acs.org on July 7, 2018

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

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

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

The Journal of Physical Chemistry Letters

Kinetics Study of OH Uptake onto Deliquesced NaCl Particles by Combining Laser Photolysis and LaserInduced Fluorescence Yosuke Sakamoto,†‡§* Jun Zhou,† Nanase Kohno,† Maho Nakagawa,†,¶ Jun Hirokawa#, and Yoshizumi Kajii†‡§ †

Graduate School of Global Environmental Studies, Kyoto University, Kyoto 606-8501, Japan



Graduate School of Human and Environmental Studies, Kyoto University, Kyoto 606-8501, Japan §

Center for Regional Environmental Research, National Institute for Environmental Studies, Ibaraki 305-8506, Japan #

Faculty of Environmental Earth Science, Hokkaido University, Sapporo 060-0810, Japan

Present address: ¶

Graduate School of Science, Hiroshima University, Hiroshima 739-8526, Japan

1 Environment ACS Paragon Plus

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

ABSTRACT

Despite the role of hydroxyl radical (OH) uptake onto sea-salt particles as a daytime chlorine source, affecting the chemical processes in the marine boundary layer, its uptake coefficient has not yet been confirmed by direct measurement methods. This study reports the application of a combination technique of laser flash photolysis generation and laser-induced fluorescence detection for the direct kinetic measurement of OH uptake onto deliquesced NaCl particles. The uptake coefficient was not constant and inversely depended on the initial OH concentration, indicating that the first uptake step is Langmuir-type adsorption. The resistance model, including surface processes, well reproduced the observed uptake coefficient. The model predicted an uptake coefficient for the atmospheric relevant OH concentration within the range from 0.77 to 0.95. Such values may lead to emissions of Cl2 higher than those predicted in previous studies based on other values. Hence, the proposed value may provide more reliable estimations of ozone formation, oxidation of volatile organic compounds, secondary organic aerosol formation, and lifetime of methane and elemental mercury in the marine boundary layer.

TOC GRAPHICS

2 Environment ACS Paragon Plus

Page 2 of 17

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

The Journal of Physical Chemistry Letters

The hydroxyl radical (OH) plays the most important role in atmospheric oxidation, because it has high reactivity toward atmospheric trace gases and initiates their oxidation process.1 In addition to gas-phase oxidation, OH takes part in the heterogeneous oxidation of both inorganic and organic aerosol particles.2 An important aspect of this heterogeneous oxidation is the emission of photochemically active gas-phase species, especially halogens, by reacting with sea-salt particles and a frozen/ice seawater surface.3-6 The heterogeneous reaction of OH with aqueous chloride is a potential source of Cl2.5-6 The proposed mechanism involves both interfacial and bulk-phase formation processes:6 Interfacial process:

OH(g) + Cl− (interface) → OHLCl− (interface)

(R1)

OHLCl− (interface) + OH LCl− (interface) → Cl2 + 2OH−

(R2)

Bulk-phase process: OH(g) → OH(aq)

(R3)

OH(aq) + Cl− (aq) ⇌ HOCl− (aq)

(R4)

2HOCl− (aq) + 2Cl− (aq)+2H+ →→ Cl2 + 2Cl− (aq) + 2H2O

(R5)

These processes can potentially affect ozone formation, volatile organic compound (VOC) oxidation, secondary organic aerosol (SOA) formation, and lifetime of elemental mercury (Hg0) and methane in the Arctic and coastal areas, because they work as a daytime source of Cl2 that produces reactive Cl in sunlight.4, 6-9

3 Environment ACS Paragon Plus

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

Page 4 of 17

To quantify the impact of heterogeneous reactions, the model simulation requires a kinetic parameter for the radical uptake. The kinetic parameter used for the uptake of gas-phase species is a net uptake coefficient, γnet (or an observable uptake coefficient, γobs), defined as follows:1

γ net (or γ obs ) =

net rate of OH(g) uptake collision rate of OH(g) with surface

(1)

To avoid the rate limitation by gas-phase diffusion, two methods are typically used for the OH uptake measurement. One is a low-pressure coated-wall flow tube method combined with OH decay measurement.10-13 The other is an aerosol flow tube/chamber experiment combined with particle-phase analysis by mass spectrometry.14-20 However, these methods are focused on OH uptake onto organic films or particles and have never been used to measure the uptake coefficient for aqueous surfaces containing Cl− because of experimental limitations. Laskin et al. used a computer-controlled scanning electron microscope with an energydispersive X-ray spectrometer (CCSEM/EDX) to directly estimate the uptake coefficient by measuring the change of the Cl-to-Na ratio in the particles.21 However, in this case, the gas-phase diffusion was an experimental constraint, and they found the value 0.1 as a lower limit.21 At present, only the uptake coefficient provided by indirect methods is available; e.g., Knipping and Dabdub estimated a value of ~0.2 by model simulation combined with measured Cl2 concentrations.22 We developed a combination technique of laser flash photolysis generation and laser-induced fluorescence (LP–LIF) to directly measure the uptake of radicals onto particles with a small gasphase diffusion constraint. This paper reports the first application of this method to measure the

4 Environment ACS Paragon Plus

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

The Journal of Physical Chemistry Letters

uptake coefficient of OH onto deliquesced NaCl particles. We carried out our experiment in a laser photolysis cell continuously flushed with a deliquesced NaCl particles/O3(g)/H2O(g)/zero air mixture at 1 atm, 295 K, and 90% relative humidity (RH). Table 1 summarizes the OH uptake coefficient onto deliquesced NaCl particles measured by LP– LIF along with previously reported values. Uptake coefficients were corrected for gas-phase diffusion as described in the Supporting Information.

Table 1 Uptake coefficients for OH onto deliquesced NaCl particles. [OH]0 / 1010 −3

molecules cm

Uptake

Surface

RH

coefficient, γ

accommodation, αs,0

/%

0.4–2.1

0.29–0.58

90

0.1–0.7

>0.1

70–80

1.1–1.3

0.17

82

~0.2

Methods

Ref.

LP–LIF

This work

CCSEM/EDX singleparticle analysis

Laskin et al.21

Chamber experiment

Knipping and

and model simulation

Dabdub22

0.95

MD simulation

Vieceli et al.23

>0.1*

Recommendation

IUPAC24

Recommendation

JPL25

>0.1

*Proposed as the lower limit for the accommodation coefficient.

Our values ranged from 0.29 to 0.58 depending on the initial OH concentration, [OH]0. Figure 1 shows the [OH]0 dependence of the OH uptake coefficient onto deliquesced NaCl particles.

5 Environment ACS Paragon Plus

The Journal of Physical Chemistry Letters

1.0

Uptake coefficient

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

Obs. with αs,0=0.95, Γs=0

0.8

with αs,0=0.95, Γs>>Γs,b, Γb

0.6

Knipping and Dabdub (2002) Laskin et al. (2006)

0.4 0.2 0.0 0.0

0.5

1.0

1.5 10

2.0

[OH]0 / 10 molecules cm

2.5 -3

Figure 1 Initial OH concentration [OH]0 dependence of the OH uptake coefficient onto deliquesced NaCl particles. The red bar and shaded area represent the lower limit and the range measured by Laskin et al.21 The green bar is the value estimated by Knipping and Dabdub.22 The solid and dashed black lines represent the fitting lines of the experimental values of this study using eq. (3) with αs,0 = 0.95,23 assuming negligible (Γs = 0) and very fast (Γs >> Γs,b and Γb) surface reactions, respectively.

As shown in Figure 1, γ decreases with increasing [OH]0. Slade and Knopf reported the same behavior when measuring OH uptake onto organic substrates and proposed a mechanism involving the Langmuir-type adsorption mechanism to explain such [OH]0 dependence.10 In the Langmuir-type adsorption, an initial adsorption of OH onto unoccupied reactive sites of the surface is followed by the uptake process including surface reactions, accommodation into the bulk, and bulk reactions.26-28 Additionally, according to molecular dynamics simulations, OH oxidation proceeds at the air–water interface via Langmuir-type adsorption owing to the rapid

6 Environment ACS Paragon Plus

Page 6 of 17

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

The Journal of Physical Chemistry Letters

thermal accommodation with absorption and desorption, in a time scale of hundreds of picoseconds.23 According to these previous papers, OH uptake onto the deliquesced NaCl particles likely proceeds via Langmuir-type adsorption. As shown in Figure 1, we found values all within the region over 0.1 reported by Laskin et al.21 and higher than the value of 0.17 estimated by Knipping and Dabdub.22 Knipping and Dabdub determined that the OH uptake coefficient reproduced observed Cl2 in chamber experiments by chemical kinetic model simulation. In their experiment, the reaction time was ~1.5 h,22 which is much longer than our reaction time below 1 s and may be long enough to change the particle’s surface condition. Additionally, because the emission process is more complicated than the uptake process, their method based on product analysis may be affected by an estimation error. Assuming that the uptake processes proceed following Langmuir-type adsorption and that the steady-state condition between elementary processes is immediately achieved, the reciprocal of γ can be described as the summation of the inverse of the rate of each process normalized to the rate of gas-surface collisions, namely conductance, as follows:26-28 1

γ

=

1

αs

1

+ Γs +

1

(2)

1 1 + Γ s,b Γ b

where αs is the surface mass accommodation coefficient and Γs, Γs,b, and Γb are the conductance values for the surface reaction, the surface–bulk mass transport, and the bulk-phase reaction, respectively. Γs, Γs,b, and Γb are proportional to αs, and γ can be expressed as follows:26  1  1+ Z

γ = αs 

  1 1    = α s,0    1 + K 'ads [OH(g)]   1 + Z

  1    = γ0    1 + K 'ads [OH(g)] 

7 Environment ACS Paragon Plus

(3)

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

where Z is the second term in eq. (2) multiplied by αs being a function of chloride concentration,

αs,0 and γ0 are the surface accommodation and the uptake coefficient on an adsorbate-free surface, respectively, and K′ads is the effective adsorption equilibrium constant. Γb is expressed as1, 26, 29

Γ b = α sCb

4 H OH RT DOH,aq kb II [Cl− (particle)]

ωOH

(4)

where HOH and DOH,aq are Henry’s law constant and the aqueous-phase diffusion coefficient for OH of 38 M atm−1 and 2.3 × 10−5 cm2 s−1, respectively, R is the gas constant, and kbII is the apparent second-order rate constant in the aqueous phase of 4.3 × 109 M−1 s−1.30-31 Cb is the reacto-diffusive geometry correction factor determined by the particle radius, rp, and the reactodiffusive length, l = (DOH,aq/kbII[Cl−(particle)])1/2:26 

Cb =

rp

∫  coth l

rp





l   dS (rp ) rp 

∫ dS (r )

(5)

p

rp

where S(rp) is a surface area distribution function at each particle radius. At ~90% RH, the mass fraction of NaCl in the particles is ~14 wt%, and the particles have a water activity of 0.90.32 With a Cl− concentration of 4 M, Cb is almost unity, leading to Γb/αs is 40. This value is large enough to consider that the term of 1/Γb in eq. (2) has a negligible contribution. Γs,b is expressed by the following equation:26

Γs,b = α s

ks,b kd

(6)

8 Environment ACS Paragon Plus

Page 8 of 17

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

The Journal of Physical Chemistry Letters

where ks,b is the first-order rate coefficient for surface-to-bulk transfer and kd is the first-order desorption rate coefficient. Although there are no experimental values available, αs,0 and ks,b/kd have been theoretically reported as 0.95 and 4.9, respectively.23 Note that the ks,b/kd is obtained from the accommodation coefficient, α (=ks,b/(ks,b+ kd)) of 0.83. These values are used in the following discussion. Because no information about Γs (=αsks/kd, where ks is the pseudo-first-order rate coefficient for chemical loss in the interfacial layer)26 is available, we considered two extreme cases: a negligible surface reaction (Γs ~ 0) and a dominant surface reaction (Γs >> Γs,b and Γb). Assuming negligible Γs and using the theoretical parameters, we estimated γ0 as 0.77. Figure 1 shows a fitting curve by eq. (3) with γ0 fixed at 0.77, which agrees with the observed values giving K′ads as (0.8 ± 0.2) × 10−10 cm3 molecule−1. Figure 1 also shows another fitting curve by eq. (3) with γ0 as 0.95 assuming the case of Γs >> Γs,b and Γb, which also agrees with the experimental values within the error bars giving K′ads as (1.3 ± 0.3) × 10−10 cm3 molecule−1. While the both cases well explain the experimental results, the parameters assuming negligible give better coefficient of determination, R2, which might suggest smaller contribution of surface reaction than bulk-phase reaction. However, we could not conclude how significant the surface reaction is in the uptake process due to a large experimental uncertainty, and thus determining the role of the surface reaction process requires further investigation of OH uptake at lower concentrations. According to the resistance model combined with our results, the uptake coefficient for atmospheric relevant OH concentrations (i.e.,., [OH]0 < ~1 × 107 molecules cm−3)1 falls within the range from 0.77 to 0.95 depending on the contribution of the surface reaction. These

9 Environment ACS Paragon Plus

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

predicted values are much higher than the previous values of ~0.2 and >0.1 recommended by IUPAC24 and JPL,25 respectively. von Glasow performed model simulation and estimated the contribution of chlorine emission by the heterogeneous reaction between OH and sea-salt particles at the marine boundary layer (MBL) considering gas-phase diffusion limitation.33 Table 2 shows the maximum relative differences in gas-phase Cl2 and Cl during daytime between the cases with γ = 0.2 and 1.0, based on a report by von Glasow.33

Table 2 Maximum relative differences of daytime Cl2 and Cl content between the cases with

γ = 0.2 and 1.0.33 Species

Coastal MBL

Remote MBL

Pristine MBL

∆/% Cl2

190–310

530–670

240–260

Cl

13–14

11

3

His calculations predicted that the chlorine emission strongly depends on the uptake coefficient used for the simulation; daytime Cl2 and Cl increase by up to 670% and 14%, respectively, if γ = 1.0 is used instead of the typical value of 0.2. Note that the surface area of atmospheric sea salt particles is typically dominated by particles of micrometer size34 onto which OH uptake is strongly limited by gas-phase diffusion. With particles over 3 micrometers in radius, γ = 1.0 and 0.2 give almost same γnet including gas-phase diffusion limitation, which agree within 20%. However, sub-micron size sea salt particles also have a contribution to the surface area, for which the gas-phase diffusion limitation is small. Therefore, the contribution of smaller particles may cause a difference in the model simulation of the chlorine cycle between γ = 1.0 and 0.2. 10 Environment ACS Paragon Plus

Page 10 of 17

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

The Journal of Physical Chemistry Letters

In field measurements near coastal areas7, 35 and at inland areas influenced by air mass from the coast,8 daytime Cl2 peaked during the afternoon with concentrations around few–several ppt and sometimes reached ~20 ppt in spite of its high photochemical reactivity. The daytime maximum concentration of Cl2 of 3.8 ppt in inland area during the afternoon indicates the presence of a photochemical Cl2 production mechanism with the production rate of up to 35 ppt h−1.8 Finley and Saltzman reported that a constant daytime Cl2 of 3.5 ppt in the coastal MBL mixed with urban air could mediate the oxidation of VOCs, resulting in an additional 5–8 ppb ozone formation.35 They also suggested that this level of daytime Cl2 enhances the oxidation of airborne elemental mercury (Hg0), contributing to mercury deposition in polluted coastal air.7 Faxon et al. estimated the daytime photochemical and heterogeneous Cl2 source strength in inland areas affected by air mass from coastal regions.8 They reported that daytime Cl2 emission enhances the maximum daily O3 and RO2 concentrations by 8–10% and 28–50%, respectively, which also contributes to SOA formation. Furthermore, daytime Cl2 emission may regionally influence the lifetime and oxidation of VOCs, including methane. Despite its importance, such high daytime Cl2 concentration levels have never been explained by the presently accepted emission mechanism.35 The presence of an additional photochemical production mechanism, including heterogeneous reactions driven by O3 or OH, has been suggested.7-8 The reaction of OH with sea-salt particles is a plausible Cl2 source during daytime. Finley and Saltzman calculated the daytime Cl2 concentration resulting from the heterogeneous OH uptake onto sea salt with γ ~0.2, which is recommended by Knipping and Dabdub22, and showed that the estimated Cl2 concentration is a factor of 7 lower than the value observed.35 However, as shown in Table 2, if the higher γ proposed in this study is used, the model prediction may increase by several times depending on the condition and may provide better

11 Environment ACS Paragon Plus

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

explanation on the observed daytime Cl2 concentration. Our results suggest the necessity of reevaluating the contribution of the heterogeneous reaction of OH and sea-salt particles as a daytime Cl2 source and their impact on the MBL.

Method The uptake coefficient of OH onto the deliquesced NaCl particles was measured at 1 atm, room temperature, and RH = 90% by a combination technique (LP–LIF), whose details are provided in the Supporting Information. We generated deliquesced NaCl particles from a 0.1 g L−1 aqueous NaCl solution (99.5%, Nacalai Tesque Inc.) with a collision-type atomizer (Model 3076, TSI, USA) at zero air (of 1.5– 3.0 slm (standard liter per minute), at 273 K and 1 atm) produced by a zero air generator (Model 111, Thermo Fisher Scientific Inc., USA). The particles were introduced into the reaction cell after being mixed with humidified zero air and O3 to have a total flow rate of 12 slm. A scanning mobility particle sizer (SMPS, Models 3936 and 3938, TSI Inc.) ensured aerosol size distributions in the reaction cell during separate experiments under the same conditions. To generate OH, sample air was irradiated once introduced into the reaction cell by a pulsed 266 nm Nd:YAG laser (Tempest 300, New Wave Research Inc., USA) with a repetition rate of 1 Hz. The initial OH concentration was calculated as 0.3–2.6 × 1010 molecules cm−3 using laser power, the absorption cross section of O3,36 the quantum yield of O(1D),37 the quenching rate constant by N2 and O2, and the rate constant of the reaction of O(1D) with H2O.38 We introduced a part of the main flow into the detection cell. Then, the probe laser, which was set at a maximum absorption of the Q1(2) A2Σ+(v′ = 0) ← X2Π3/2(v′′ = 0) transition of OH at 308 nm with a repetition rate of 10 kHz (Sirah Credo, Spectra Physics), was irradiated. We measured

12 Environment ACS Paragon Plus

Page 12 of 17

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

The Journal of Physical Chemistry Letters

the fluorescence from OH after each probe laser irradiation to obtain a decay curve, whose profile is a single-exponential decay:

[OH] = [OH]0 exp(−(k'PU +k'BG )t )

(7)

where k′BG is the background loss and k′PU is the first-order loss rate by particles with γobs given by26 k 'PU =

γ obsωOH S

(8)

4

where ωOH is the mean thermal velocity for OH and S is the surface concentration of particles. Because γobs includes the contribution of the gas-phase diffusion, we corrected it based on the approach of Fuchs and Sutugin to extract the contribution of the particle-related process, γ.39 Further details of the method and data analysis are available in the Supporting Information.

ASSOCIATED CONTENT *Supporting Information Details of experiments and data analysis with figures: schematic diagram of the experimental setup (Figure S1), examples of particle size distribution (Figure S2), examples of OH radical decay (Figure S3), and dependence of loss rate by particle uptake on surface area concentration (Figure S4).

AUTHOR INFORMATION Corresponding Author *(Yosuke Sakamoto) Email: [email protected], Tel: +81-75-753-6826.

ORCID

13 Environment ACS Paragon Plus

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

Yosuke Sakamoto: 0000-0002-9863-4241

Notes The authors declare no competing financial interests.

ACKNOWLEDGMENT This work was supported by JSPS KAKENHI Grant Numbers JP16H06305 and JP16K16183. The authors would like to thank Enago (www.enago.jp) for the English language review.

14 Environment ACS Paragon Plus

Page 14 of 17

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

The Journal of Physical Chemistry Letters

REFERENCES

(1) Finlayson-Pitts, B. J.; J. N. Pitts, J. Chemistry of the Upper and Lower Atmosphere. Academic Press, California, USA: 1999. (2) George, I. J.; Abbatt, J. P. D. Heterogeneous Oxidation of Atmospheric Aerosol Particles by Gas-Phase Radicals. Nat. Chem. 2010, 2, 713-722. (3) Abbatt, J.; Oldridge, N.; Symington, A.; Chukalovskiy, V.; McWhinney, R. D.; Sjostedt, S.; Cox, R. A. Release of Gas-Phase Halogens by Photolytic Generation of OH in Frozen Halide−Nitrate Solutions: An Active Halogen Formation Mechanism? J. Phys. Chem. A 2010, 114, 6527-6533. (4) Frinak, E. K.; Abbatt, J. P. D. Br2 Production from the Heterogeneous Reaction of GasPhase OH with Aqueous Salt Solutions:  Impacts of Acidity, Halide Concentration, and Organic Surfactants. J. Phys. Chem. A 2006, 110, 10456-10464. (5) Oum, K. W.; Lakin, M. J.; DeHaan, D. O.; Brauers, T.; Finlayson-Pitts, B. J. Formation of Molecular Chlorine from the Photolysis of Ozone and Aqueous Sea-Salt Particles. Science 1998, 279, 74-76. (6) Knipping, E. M.; Lakin, M. J.; Foster, K. L.; Jungwirth, P.; Tobias, D. J.; Gerber, R. B.; Dabdub, D.; Finlayson-Pitts, B. J. Experiments and Simulations of Ion-Enhanced Interfacial Chemistry on Aqueous NaCl Aerosols. Science 2000, 288, 301-306. (7) Finley, B. D.; Saltzman, E. S. Observations of Cl2, Br2, and I2 in Coastal Marine Air. J. Geophys. Res. Atmos. 2008, 113, D21301. (8) Faxon, C.; Bean, J.; Ruiz, L. Inland Concentrations of Cl2 and ClNO2 in Southeast Texas Suggest Chlorine Chemistry Significantly Contributes to Atmospheric Reactivity. Atmosphere 2015, 6, 1487-1506. (9) Knipping, E. M.; Dabdub, D. Impact of Chlorine Emissions from Sea-Salt Aerosol on Coastal Urban Ozone. Environ. Sci. Technol. 2003, 37, 275-284. (10) Slade, J. H.; Knopf, D. A. Heterogeneous OH Oxidation of Biomass Burning Organic Aerosol Surrogate Compounds: Assessment of Volatilisation Products and the Role of OH Concentration on the Reactive Uptake Kinetics. Phys. Chem. Chem. Phys. 2013, 15, 5898-5915. (11) Slade, J. H.; Knopf, D. A. Multiphase OH Oxidation Kinetics of Organic Aerosol: The Role of Particle Phase State and Relative Humidity. Geophys. Res. Lett. 2014, 41, 5297-5306. (12) Cooper, P. L.; Abbatt, J. P. D. Heterogeneous Interactions of OH and HO2 Radicals with Surfaces Characteristic of Atmospheric Particulate Matter. J. Phys. Chem. 1996, 100, 2249-2254. (13) Bertram, A. K.; Ivanov, A. V.; Hunter, M.; Molina, L. T.; Molina, M. J. The Reaction Probability of OH on Organic Surfaces of Tropospheric Interest. J. Phys. Chem. A 2001, 105, 9415-9421. (14) George, I. J.; Vlasenko, A.; Slowik, J. G.; Broekhuizen, K.; Abbatt, J. P. D. Heterogeneous Oxidation of Saturated Organic Aerosols by Hydroxyl Radicals: Uptake Kinetics, Condensed-Phase Products, and Particle Size Change. Atmos. Chem. Phys. 2007, 7, 4187-4201. (15) Hearn, J. D.; Smith, G. D. A Mixed‐Phase Relative Rates Technique for Measuring Aerosol Reaction Kinetics. Geophys. Res. Lett. 2006, 33, L17805. (16) Lambe, A. T.; Zhang, J.; Sage, A. M.; Donahue, N. M. Controlled OH Radical Production via Ozone-Alkene Reactions for Use in Aerosol Aging Studies. Environ. Sci. Technol. 2007, 41, 2357-2363.

15 Environment ACS Paragon Plus

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

(17) Lambe, A. T.; Miracolo, M. A.; Hennigan, C. J.; Robinson, A. L.; Donahue, N. M. Effective Rate Constants and Uptake Coefficients for the Reactions of Organic Molecular Markers (n-Alkanes, Hopanes, and Steranes) in Motor Oil and Diesel Primary Organic Aerosols with Hydroxyl Radicals. Environ. Sci. Technol. 2009, 43, 8794-8800. (18) Smith, J. D.; Kroll, J. H.; Cappa, C. D.; Che, D. L.; Liu, C. L.; Ahmed, M.; Leone, S. R.; Worsnop, D. R.; Wilson, K. R. The Heterogeneous Reaction of Hydroxyl Radicals with SubMicron Squalane Particles: A Model System for Understanding the Oxidative Aging of Ambient Aerosols. Atmos. Chem. Phys. 2009, 9, 3209-3222. (19) McNeill, V. F.; Yatavelli, R. L. N.; Thornton, J. A.; Stipe, C. B.; Landgrebe, O. Heterogeneous OH Oxidation of Palmitic Acid in Single Component and Internally Mixed Aerosol Particles: Vaporization and the Role of Particle Phase. Atmos. Chem. Phys. 2008, 8, 5465-5476. (20) Renbaum, L. H.; Smith, G. D. Artifacts in Measuring Aerosol Uptake Kinetics: The Roles of Time, Concentration and Adsorption. Atmos. Chem. Phys. 2011, 11, 6881-6893. (21) Laskin, A.; Wang, H.; Robertson, W. H.; Cowin, J. P.; Ezell, M. J.; Finlayson-Pitts, B. J. A New Approach to Determining Gas-Particle Reaction Probabilities and Application to the Heterogeneous Reaction of Deliquesced Sodium Chloride Particles with Gas-Phase Hydroxyl Radicals. J. Phys. Chem. A 2006, 110, 10619-10627. (22) Knipping, E. M.; Dabdub, D. Modeling Cl2 Formation from Aqueous NaCl Particles: Evidence for Interfacial Reactions and Importance of Cl2 Decomposition in Alkaline Solution. J. Geophys. Res. Atmos. 2002, 107, 4360. (23) Vieceli, J.; Roeselová, M.; Potter, N.; Dang, L. X.; Garrett, B. C.; Tobias, D. J. Molecular Dynamics Simulations of Atmospheric Oxidants at the Air−Water Interface:  Solvation and Accommodation of OH and O3. J. Phys. Chem. B 2005, 109, 15876-15892. (24) Ammann, M.; Cox, R. A.; Crowley, J. N.; Jenkin, M. E.; Mellouki, A.; Rossi, M. J.; Troe, J.; Wallington, T. J. Evaluated Kinetic and Photochemical Data for Atmospheric Chemistry: Volume VI – Heterogeneous Reactions with Liquid Substrates. Atmos. Chem. Phys. 2013, 13, 8045-8228. (25) Burkholder, J. B.; Sander, S. P.; Abbatt, J. P. D.; Barker, J. R.; Huie, R. E.; Kolb, C. C.; Kurylo, M. J.; Orkin, V. L.; Wilmouth, D. M.; Wine, P. H. Chemical Kinetics and Photochemical Data for Use in Atmospheric Studies, Evaluattion No. 18. http://jpldataeval.jpl.nasa.gov (accessed June 25th, 2018). (26) Pöschl, U.; Rudich, Y.; Ammann, M. Kinetic Model Framework for Aerosol and Cloud Surface Chemistry and Gas-Particle Interactions - Part 1: General Equations, Parameters, and Terminology. Atmos. Chem. Phys. 2007, 7, 5989-6023. (27) Ammann, M.; Poschl, U.; Rudich, Y. Effects of Reversible Adsorption and LangmuirHinshelwood Surface Reactions on Gas Uptake by Atmospheric Particles. Phys. Chem. Chem. Phys. 2003, 5, 351-356. (28) Hanson, D. R. Surface-Specific Reactions on Liquids. J. Phys. Chem. B 1997, 101, 49985001. (29) Danckwerts, P. V. Gas-Liquid Reactions. McGraw-Hill Book Co.: New York, U.S.; 1970. (30) Buxton, G. V.; Greenstock, C. L.; Helman, W. P.; Ross, A. B. Critical Review of Rate Constants for Reactions of Hydrated Electrons, Hydrogen Atoms and Hydroxyl Radicals (⋅OH/⋅O−) in Aqueous Solution. J. Phys. Chem. Ref. Data 1988, 17, 513-886. (31) Sander, R. Compilation of Henry's Law constants (version 4.0) for Water as Solvent. Atmos. Chem. Phys. 2015, 15, 4399-4981.

16 Environment ACS Paragon Plus

Page 16 of 17

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

The Journal of Physical Chemistry Letters

(32) Tang, I. N.; Tridico, A. C.; Fung, K. H. Thermodynamic and Optical Properties of Sea Salt Aerosols. J. Geophys. Res. Atmos. 1997, 102, 23269-23275. (33) von Glasow, R. Importance of the Surface Reaction OH + Cl- on Sea Salt Aerosol for the Chemistry of the Marine Boundary Layer – a Model Study. Atmos. Chem. Phys. 2006, 6, 35713581. (34) O'Dowd, C. D.; Smith, M. H.; Consterdine, I. E.; Lowe, J. A. Marine Aerosol, Sea-Salt, and the Marine Sulphur Cycle: a Short Review. Atmos. Environ. 1997, 31, 73-80. (35) Finley, B. D.; Saltzman, E. S. Measurement of Cl2 in Coastal Urban Air. Geophys. Res. Lett. 2006, 33, L11809. (36) Gorshelev, V.; Serdyuchenko, A.; Weber, M.; Chehade, W.; Burrows, J. P. High Spectral Resolution Ozone Absorption Cross-Sections -Part 1: Measurements, Data Analysis and Comparison with Previous Measurements around 293 K. Atmos. Meas. Tech. 2014, 7, 609-624. (37) Matsumi, Y.; Comes, F. J.; Hancock, G.; Hofzumahaus, A.; Hynes, A. J.; Kawasaki, M.; Ravishankara, A. R. Quantum Yields for Production of O(1D) in the Ultraviolet Photolysis of Ozone: Recommendation Based on Evaluation of Laboratory Data. J. Geophys. Res. Atmos. 2002, 107, 4024. (38) Atkinson, R.; Baulch, D. L.; Cox, R. A.; Jr., R. F. H.; Kerr, J. A.; Rossi, M. J.; Troe, J. Evaluated Kinetic and Photochemical Data for Atmospheric Chemistry: Supplement VI. IUPAC Subcommittee on Gas Kinetic Data Evaluation for Atmospheric Chemistry. J. Phys. Chem. Ref. Data 1997, 26, 1329-1499. (39) Fuchs, N. A.; Sutugin, A. G. High-Dispersed Aerosols. In Topics in Current Aerosol Research, Hidy, G. M.; Brock, J. R., Eds. Pergamon: 1971; pp 1-60.

17 Environment ACS Paragon Plus